a?” ”in?“ 0. RM Mg' :5... 'h n. Wmf“"" $.23“; .x‘z" "hf ““145”? 321%? 34) :6 1' , , 4: r 5 ~ . ”Emma-- ' .- 3.,“- “ ' 4 ‘ 4-7593: , . {h ‘ 1. ‘ 38:54:24.! ’ 1mm 5-,, ‘73: .39“"" 4 ‘ , 21.5,, 93“” 4. i 4~115m€ | V . ‘ {fir-I" P'jfi'l'l' . 4:” £339; 3,3 «3:35.- ”4;; .3" 14.1%,?» £323.», 553, {a 5'95“ ' :‘j’ . ‘ 1 it ‘2. a. H I | ' n.‘ 93.}: W _:.‘q‘r::é :I’,‘ 3‘ lit-:g‘wz‘ifA’kK' ' r, ‘ “ ,- '::,:.' " 39‘“? 3‘44“" I ; _ "‘ .2... ask‘ _ ~ficl-J“ - ‘ . ~f r h-$ 3.7%.]? ‘ "‘ if“: .,-. ._gqu£. _ «ow V 74:94:} .‘ . _ J‘ ' 3 {gr- . .a. u . . . P X “H .....»..:y, ”‘1‘ ‘. RA v 5 $3433.13, 3'3"}, 4 if: ri’fi-J ifig‘ ' - 3.. ’ ‘ A“ “J 7: '3' . m ‘w. ~14 . 4 «4 . u d, . _ @‘gfi‘fiufig'm ' ' ’ - 4 $§$ 8' u. 7 r . ‘ ' '-' 3 . . u? ' - : ~ : - . , . ‘4‘) “ E" - I. Wam- ’ -} 4 . ;. , (a . , 4‘? , . mi}: 5 4’ '1‘ 1,7 ~- _ - - v 4‘1 {F ”up. 1 '3"! u . ' «r fiiwmx -’ tJ v1} ' 4.}912 “£3. ‘ 5;}: . ’4' w. {11139: ... .5‘ y- c n, ..,_ .- ‘al. ‘1 ”4 h “1;. L54,“ ‘1' 345$ ’21, c, .. - 4°43: 1- 4 a... r- .. :53 7..,.{.>t”"l:. i1 , Jana?“ ‘ ‘ my; 1:.“ “’3 ,-" 4 “3“ A ~. “If“: ‘ 29:31: :3 ‘H “44,-"rfi, an {33.14: 4 .1» . , ' ~ ' ‘;,§.r...r.,n,:.,, 4. 4.. , 1 _ - g , .- ”. V a» " WWW M r ”’55"? ‘ ‘ r ' ' ‘ ‘4} 4?“ *5 H. ; g, f ' 4_‘;:;.§: , . 2' #:j-sff‘ .. .214 x ‘ "L". M «yr. :. z 4", h" " Q». ' J 34} 1“”.53- »J)". .ar 4:: .'l' . J'err‘v ': . '39-‘18 This is to certify that the dissertation entitled Active site studies on Klebsiella aerogenes urease, a nickel-containing enzyme presented by Matthew J. Todd has been accepted towards fulfillment of the requirements for PhD degree in Biochemistry 16% WW \ Major professor / Date June 19, 1991 MSU is an Affirmative Action/Equal Opportunity Institution ‘ 0-12771 _-——--—__——-‘,-—_A ‘1 — ”£444,444 7 M‘Chigan 4.44444; Unwersétgv J PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. r DATE DUE DATE DUE DATE DUE H MSU Is An Affirmative Action/Equal Opportunity Institution cWafl-M ACTIVE SITE STUDIES ON KLEBSIELLA AEROGENES UREASE, A NICKEL-CONTAINING ENZYME BY Matthew J. Todd A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Biochemistry 1991 ABSTRACT ACTIVE SITE STUDIES ON KLEBSIELLA AEROGENES UREASE, A NICKEL-CONTAINING ENZYME BY Matthew J. Todd Although ureases play important roles in microbial ni- trogen metabolism and in the pathogenesis of several human diseases, little is known of the mechanism of urea hydro- lysis. As a paradigm for microbial ureases, the enzyme from Klebsiella aerogenes was studied. This enzyme was purified 1070-fold and shown to consist of three polypeptides (esti— mated.bb.= 72,000, 11,000, and 9,000) in an apparent oqflnm quaternary structure, containing 4 mol nickel/mol urease. A tight-binding inhibitor was exploited to demonstrate the presence of two mol active site/mol enzyme, thus each cata- lytic unit ((13272) has two nickel. The heteropolymeric sub- unit structure was not expected since the well-studied plant urease from the jack bean is homopolymeric. K. aerogenes urease is competitively inhibited by thiols, boronic acids, hydroxamates, phosphate and phos- phoroamides. UV-visible absorbance studies demonstrated charge-transfer interactions between thiolate anions and urease nickel ion, consistent with the presence of at least one nickel per catalytic unit. Comparison of.K;velues for several thiolate inhibitors were consistent with electrosta- tic repulsion of anionic inhibitors by an anionic group on the enzyme near the thiolate binding site. Acetohydroxamate (AHA) and phenylphosphorodiamidate (PPD) were slow-binding competitive inhibitors with ngalues of 2.6 uM and 94 pM, respectively. Following initial binding to one nickel, the inhibitors were proposed to form complexes bridging the two urease active site nickel ions. The kinetics of K. aerogenes urease thiol modification were studied using alkylating and disulfide reagents. Reac- tivity of the essential thiol was affected by substrate and competitive inhibitors, consistent with a cysteine located proximal to the active site. The atypical pH dependence for the rate of inactivation was interpreted as arising from an interaction between the thiol and a second (as yet unidenti— fied) ionizing group. The competitive inhibitors phosphate and PPD (but not AHA) protected one mol thiol/mol catalytic unit from.modification by 2,2'dithiodipyridine. Using 14C- iodoacetamide, the essential cysteine was identified as Cys319, an amino acid conserved among all known urease se- quences. to my wife and family AC KN OWL EDGMENTS As with any major work, the production of a dissertation is bound to inconvenience and even burden those around us. Recognizing this, I would like to acknowledge two individuals whose sacrifices helped to make it possible. Dr. Hausinger, not only as mentor, but also as a friend gave me much needed guidance and then retreated, allowing me to fail or succeed on my own. Janet, my wife, was willing to sacrifice her own time and energy, allowing me to pursue these studies. During most of our marriage, she has endured considerably less at- tention than she desired, without excess complaint, while reminding me that there is a "real" world, around which we should ideally focus our lives. I would like to thank Scott Mulrooney for providing pKAUlQ, a plasmid which expresses urease to almost 10% of the cell protein, Mann Hyung Lee for one preparation of K. aerogenes urease, Dr. J. Breznek for the use of his scin— tillation counter, and Dr. J. Leykam for helpful discussions concerning peptide isolation strategy. Dr. J. Wilson and Dr. T. Deits gave insightful comments during preparation of chapters 2-4. The remaining committee members, Dr. C. Chang, Dr. J. Kaguni, and Dr. D. Lamport were available and willing to give guidance when needed. TABLE OF CONTENTS List of Tables .......................................... viii List of Figures ........................................... ix List of Abbreviations .................................... xii Chapter 1. Introduction Medical significance .................................. 2 Agricultural significance ............................. 4 Industrial significance ............................... 5 Plant ureases ......................................... 6 Properties of purified urease ......................... 6 Kinetic parameters ................................... 10 Inhibition of urease ................................. 10 Active site residues ................................. 13 References ........................................... 17 Chapter 2. Purification and Characterization of the Nickel- containing Multicomponent Urease from K. aerogenes Abstract ............................................. 28 Introduction ......................................... 30 Experimental Procedures .............................. 31 Bacterial growth conditions ..................... 31 Assays .......................................... 32 Preparation of crude extract .................... 32 SDS-polyacrylamide gel electrophoresis .......... 33 Native gel electrophoresis ...................... 34 Native molecular weight determination ........... 34 Stability studies ............................... 35 pH studies ...................................... 35 Amino acid analysis ............................. 36 Nickel analysis ................................. 36 Results .............................................. 37 Urease purification ............................. 37 SDS-polyacrylamide gel electrophoresis of K. aerogenes urease ........................... 43 Native molecular weight for K. aerogenes urease.43 Attempts to resolve the three urease components.45 Enzyme stability studies ........................ 46 Kinetic parameters .............................. 48 pH studies ...................................... 48 Amino acid analysis ............................. 49 Atomic absorption analysis ...................... 49 Discussion ........................................... 49 References ........................................... 57 Chapter 3. Competitive Inhibitors of Klebsiella aerogenes Urease: Mechanism of Interaction with the Nickel Active Site Abstract ............................................. 60 Introduction ......................................... 62 Experimental Procedures .............................. 63 Materials ....................................... 63 Enzyme purification ............................. 64 Assays .......................................... 64 Spectroscopy .................................... 65 K,- determinations ............................... 66 Inhibitor binding rates in the absence of substrate .................................. 66 Progress curves in the presence of inhibitor....67 Dissociation rate ............................... 67 Statistical methods ............................. 68 Quantitation of active sites .................... 68 Results .............................................. 69 Thiol inhibition ................................ 69 Spectral interactions of B-ME with urease ....... 72 Phosphate Inhibition ............................ 75 Acetohydroxamate inhibition ..................... 75 PPD inhibition .................................. 84 Quantitation of active sties .................... 85 Other inhibitors ................................ 85 Discussion ........................................... 87 Thiol interactions with urease .................. 89 Phosphate inhibition ............................ 91 Acetohydroxamic acid inhibition ................. 93 Phenylphosphorodiamidate inhibition ............. 96 Boric and boronic acids ......................... 99 References .......................................... 101 Chapter 4. Reactivity of the Essential Thiol of Klebsiella aerogenes Urease: Effect of pH and Ligands on Thiol Modification Abstract ............................................ 104 Introduction ........................................ 106 Experimental Procedures ............................. 107 Enzyme Purification ............................ 107 Kinetics of urease inactivation ................ 108 Reactivation of DTNB- or DTDP-inactivated urease .................................... 111 Spectroscopic analysis ......................... 111 Curve fitting .................................. 112 Results ............................................. 113 Inactivation of urease by thiol-specific reagents .................................. 113 Reactivation of urease inactivated by disulfide reagents .................................. 116 Effect of ligands on the rate of thiol modification .............................. 119 The effect of pH on leapp ....................... 123 Spectrophotometric quantitation of urease thiols .................................... 126 Discussion .......................................... 130 Demonstration of an essential thiol in K. aerogenes urease ....................... 130 Is the essential thiol at the active site? ..... 131 pH dependence of urease inactivation by IAM and DTDP .................................. 132 Urease inactivation by DTNB .................... 136 Spectroscopic quantitation of urease thiols....l38 References .......................................... 140 (A) . . Chapter 5. Identification of the Essential Cysteine Residue in Klebsiella aerogenes Urease Abstract ........ - .................................... 142 Introduction ........................................ 143 Experimental Procedures ............................. 146 Enzyme purification ............................ 146 Uniform labelling of cysteine residues in urease ................................. 146 Specific labelling of the cysteine residues in native urease ............................. 147 Standardized protocol for peptide mapping ...... 147 Characterization of the specifically labelled peptide ................................... 148 Results and discussion .............................. 148 Uniform labelling of urease cysteines and de- velopment of a peptide mapping protocol...148 Specific labelling of the essential cysteine at pH 6.3 ................................. 152 Specific labelling of the essential cysteine at high pH ................................ 154 Identification of the cysteine in peptide D....159 Comparison of K. aerogenes urease sequence to other ureases ............................. 161 References .......................................... 163 Chapter 6. Summary and Future Directions Urease purification and characterization ............. 165 Urease inhibitors .................................... 169 Active site amino acids .............................. 170 Appendix. Saturation Magnetization of the nickel centers in urease ................................... 174 Wi ~a..¢ .. v... LIST OF TABLES Chapter 2 l. Urease purification from K. aerogenes ............ 4O 2. Stability of K. aerogenes urease ................. 47 3. Amino acid compositions of ureases ............... 51 Chapter 3 1. Thiol Inhibition of K. aerogenes urease .......... 71 Chapter 4 1. Second-order Rate Constants for Loss of Urease Activity by Thiol-Specific Reagents ..... 115 2. Effect of Urease Ligands on the Rate of Inactivation by Thiol-Specific Reagents ........ 122 Chapter 5 1. Amino Acid Composition of the Essential Cysteine-Containing Peptide ................... 160 Wfi LIST OF FIGURES Chapter 2 1. DEAE—Sepharose Chromatography of Klebsiella aerogenes urease ................................ 38 2. Phenyl-Sepharose Chromatography of Klebsiella aerogenes urease ................................ 39 3. Mono Q Chromatography of Klebsiella aerogenes urease .......................................... 41 4. Superose 6 Chromatography of Klebsiella aerogenes urease ................................ 42 5. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of several microbial ureases....44 6. The effect of pH on log V /K; for max K. aerogenes urease ............................. 50 Chapter 3 1. pH dependence of cysteamine inhibition of urease.70 2. UV-visible spectra of urease in the presence of thiol ........................................... 73 3. Double-reciprocal plot of l/A absorbance versus [B-MEJ-l ........................................ 74 4. pH dependence of phosphate inhibition ............ 76 5. Urease progress curves in the presence of AHA....77 6. Rates of AHA binding in the absence of substrate.79 7. Double~reciprocal plot of hmpfl versus [AHA]4....81 8. Progress curves for AHA dissociation ............. 83 9. Quantitation of urease active sites .............. 86 Chapter 4 1. Pseudo—first—order urease activity loss in the presence of DTDP ............................... 114 2. Saturation of the apparent rate of urease modification by DTNB: the effects of urea and borate ..................................... 117 3. Reactivation of DTNB—inactivated urease ......... 118 4. Effect of the competitive inhibitors boric acid and phosphate on the rate of urease inactivation by IAM ............................ 120 5. The pH dependence of IAM and DTDP inactivation of urease ...................................... 124 6. Effect of pH and DTNB concentration on the pseudo- first-order rate of urease activity loss ....... 125 7. Spectroscopic progress curves for the reaction of urease with DTDP ............................ 127 8. Activity as a function of urease modification...129 Chapter 5 1. Subunit distribution of urease thiols ........... 149 2. Pro-RPC analysis of uniformly-labelled trypsin-digested.urease a—subunit ............. 151 3. Effect of phosphate on the modification of urease thiols by IAM ........................... 153 4. Peptide distribution of urease thiols modified at low pH ...................................... 155 5. Effect of competitive inhibitors on urease thiol modification at pH 7.75 .................. 156 6. Peptide distribution of urease thiols modified at high pH ..................................... 158 7. Comparison of urease sequences surrounding the essential cysteine ......................... 162 Chapter 6 1. Separation of urease isozymes on Mono-Q ......... 167 2. Pseudo-first order inactivation of urease by arginine specific reagents ..................... 172 LI ST 0]? ABBREVIATIONS AHA, acetohydroxamic acid; BBA, benzene boronic acid; 4-Br—BBA, 4—bromobenzene boronic acid, CHES, 2-[N-cyclo- hexylamino]~ethanesulfonic acid; CAPS, 3-[cyclohexylamino]— l—propane-sulfonic acid; DTNB, 5,5'-dithiobis(2-nitrobenzoic acid); DTDP, 2,2'-dithiodipyridine; DTT, dithiothreitol; EDTA, ethylenediamine tetraacetic acid; Gn-HCl, guanidine hydrochloride; HEPES N—[2-hydroxyethyl]piperazine-N'-[2- ethanesulfonic acid]; IAA, iodoacetate; IAM, iodoacetamide; B-ME, 2-mercaptoethanol; MMTS, methyl methanethiolsulfon— ate; MBPT, methyl butenyl phosphoric triamide; MES, 2-[N- morpholino]ethanesulfonic acid; MOPS, 3-[N-morpholino]pro- panesulfonic acid; NEM, N-ethyl maleimide; PPD, phenylphos— phorodiamidate; TAPS, N-tris[hydroxymethyl]methyl-3-amino- propanesulfonic acid; TPCK, L—l—p—tosylamino—Z-pheny1ethyl chloromethyl ketone; TFA, trifluoroacetic acid; TAPS, N- tris[hydroxymethyl]methyl—3-aminopropanesulfonic acid; Tris-HCl, tris(hydroxymethyl)aminomethane~hydrochloride. mi INTRODUCTION Urea is a remarkably stable molecule: the half-life for the spontaneous degradation of urea is 3.6 years in aqueous solutions at 38'C (6). Urea is not, however, very stable in the environment. It is rapidly hydrolyzed to ammonia and carbamate through the action of the enzyme urease (EC 3.5.1.5) (5). Carbamate spontaneously degrades to another molecule of ammonia and carbon dioxide; the car— bon dioxide is then hydrated to carbonic acid which will dissociate into H+ and Hcog. Two mol ammonia and one mol bicarbonate are the net products; thus urea hydrolysis results in a net increase in pH (Scheme 1). H30 H20 3 \_ fi \_ H,N-c-NH, .— HgN-C-OH .—_—> 002 \__ H2003 é H00; + H+ + + NH; NH; Scheme 1 Ureases are found in bacteria, algae, filamentous fungi, yeasts, and plants (reviewed by Mobley and Hausinger; 55). Pathogenic ureolytic microorganisms have been isolated from humans, where they can colonize the urinary tract, the colon, or the stomach. Urease activity is also found in soil, either associated with soil microbes, or as extra- cellular urease, probably arising from the lysis of soil microbes or the decomposition of plant tissues (61). The purpose of this chapter is to describe briefly several areas where microbial ureases affect the medical and agricultural community, then to review what is currently known about urease enzymology. .Medical significance— Urease is an important virulence factor in several pathogenic states, especially those asso— ciated with urinary tract infections. Ureolytic microorgan- isms in this environment will hydrolyze urea (at 0.4 M in human urine; 32) and cause a rise in urine pH which can lead to pyelonephritis (74). Elevated urine pH leads to supersa- turation of the urine with respect to magnesium and calcium salts (32) which can precipitate forming stones; 15-20% of all kidney stones are associated with infections by ureo- lytic microorganisms (33). The extracellular polysaccharide of the bacterium is thought to provide an organic matrix necessary for stone initiation (51) and play a role in ce- menting the crystalline mineral components together (52). Growth of urinary calculi imbeds viable bacteria in inter- stices of the urinary stone (86), inhibiting antibiotic treatment of such an infection. Precipitation of urine salts also affects patients with long term urinary catheter— ization; 86% of catheterized patients have ureolytic bac— teria present in their urine (56). In addition to adverse effects caused by the elevation of pH, ureases release vast quantities of ammonia. Hyper- ammonemia (83), hepatic encephalopathy (78), and hepatic coma (76) are conditions caused by elevated levels of am- monia, which can not be removed from the system by the liver, either because the liver is damaged or because the organism is producing an excessive amount of ammonia. Am- monia is known to have many toxic effects (15), and the additional ammonia released by ureolytic infections can contribute to the overall nitrogen burden of the organism. Treatment of stone-forming kidney infections, hyperam- monemia, or ammonia toxicity may be more effective using urease inhibitors; studies using urease inhibitors brflfiv (85) and in vivo (44, 53, 73) have demonstrated that calculus formation is reduced in urine lacking urease activity. Similarly, urease inhibitors have been demonstrated to re- duce blood ammonia levels in patients with hepatic coma (83). The rational design of effective inhibitors will re- quire detailed knowledge of the urease mechanism. A second type of infection in which urease contributes to the pathogenicity of a bacterium occurs in the mucosal lining of the mammalian stomach (23). Colonization of this hostile environment by the acid sensitive bacterium, Helico- bacter pylori occurs at intercellular junctions (10). H. pylori is the etiologic agent of type B gastritis (23, 31) and has been implicated in the formation of peptic ulcers (37); eradication of the infection relieves the symptoms (23). H. pylori possesses an active urease which can ra- pidly hydrolyze serum urea, producing abnormal amounts of ammonia, initially thought to create a ”cloud" of neutral pH immediately surrounding the bacterium (30). Current hypo- theses to account for H. pylori pathology include: bacterial secretion of mucin digesting enzymes (80), interference with the passage of H+ ions from the gastric glands to the lumen (37), and/or toxicity due to elevated ammonia concentra— tions, as mentioned above (15, 81). Supporting the role of urease as a virulence factor of H. pylori, Smoot et al. (81) demonstrated similar toxicity to human gastric adenocarcin— oma cells grown with urea and either H. pylori or jack bean urease. As in the case of urinary tract infections, urease inhibitors may aid in combatting H. pylori infections. Agricultural significance- The lack of fixed nitrogen in the soil is a major limitation to plant growth; there- fore, fertilizers are used to supplement soil nitrogen levels when growing commercially important crop plants. Urea is a good source of nitrogen, because of its low cost, ease of use, and its high nitrogen content (3). Urea must be hydrolyzed by soil ureases before it can be assimilated into plant tissue; uncontrolled hydrolysis can lead to elevated soil pH, ammonia toxicity, and nitrogen loss as volatile NH3 (up to 50% for submerged rice crops; 9). The simultaneous application of urease inhibitors and urea-based fertilizers could minimize crop damage and enhance the “HW efficiency of nitrogen utilization (8, 9). Hydroxamic acids (69) and phosphoroamide derivatives (8, 16, 50, 61) retard urea hydrolysis in soils; however, an increase in plant yield has not yet been demonstrated. One problem regarding the application of urease inhibitors has been the accumula- tion of urea in leaves, resulting in leaf-tip necrosis (45). Obviously, further work in this field is of value. Industrial significance- Ureases are also of practical use industrially. Inside a bioreactor, the delivery of nu- trients to growing cells is diffusion limited; uniform growth of cells cannot occur if the cells at the site of nutrient input deplete the medium of nutrients. Attaching urease to the solid support in the bed of bio-reactors al— lowed Mak et al. (49) to use urea as a nitrogen source for growth of a photosynthetic algae, Chlorella emersonii. Re- actors initiated with immobilized urease had a more uniform release of ammonia proximal to the growing cells and in- creased penetration of the gel matrix (48, 49). Ureases have also been used to remove urea waste from industrial effluents (G. Prabhakaran, personal communication), or from blood dialysate in artificial kidneys (67). The waste urea is hydrolyzed to NH3, which can be easily removed by air stripping. Plant ureases- Upon initiation of the current work, the only extensively studied urease was that from Canavalia ensiformis, the jack bean plant (reviewed by Blakeley and Zerner; 6); few microbial ureases had even been purified. Jack bean urease has the distinction of being the first enzyme crystallized (84) and the first enzyme shown to pos- sess nickel (19). Research on the active site of the plant enzyme allowed Zerner and colleagues to propose a mechanism for the hydrolysis of urea by urease (22), Scheme 2. In- deed, many of the initial results on K. aerogenes urease are interpreted in terms of this model in which each active site has two nickel ions: one binds to the carbonyl oxygen of urea, and the other binds a hydroxide for nucleophilic at- tack on the urea carbon. Elimination of one molecule of ammonia, followed by release of carbamate and binding of two molecules of water is thought to regenerate the resting enzyme. Although plant ureases have been found in seed, leaf and stem tissues, their role in plant nitrogen metabo- lism is not well understood (94). Properties of purified ureases- In contrast to the homohexameric structure (Mr = 90,770; 88) of jack bean urease (6) and soybean urease (68), many microbial ureases have three distinct subunits, a;B,and‘x where a = 60-70 kDa, and B and y are 8—12 kDa (39, 41, 59, 89), as first descri- bed for K. aerogenes urease (Chapter 2; 90). The B and 7 subunits have been overlooked when analyzing certain enzymes by SDS PAGE using gels of less than 12% acrylamide (12, 82). In one organism, H) pylori, the B and.y subunits have N oanom \/' l ohmEanookx ONI 00,5 / kw / / ON IN #01:. M fb undergone gene fusion into a single polypeptide of 30 kDa (46). Subunit ratios of urease enzymes have been variable among species as determined by integrated scanning densito— metry, a method which assumes that each subunit has a simi- lar affinity for the dye; therefore definitive comparisons of subunit ratios are meaningless. Genetic studies have confirmed the presence of three structural genes in K. aerogenes (60), P. stuartii (59), P. vulgaris (58), P. mirabilis (40), and H. pylori (46) and have shown that the three genes are highly honologous to the single plant gene, which probably arose via gene fusion. Furthermore, DNA sequence studies of Scott Mulrooney (60) suggested that three additional genes are involved in generation of active urease. One possible function of these additional genes is in nickel processing (47, 36); deletion of these accessory genes resulted in production of urease apoenzyme (60). Nickel has been found in all ureases examined (35). Calculation of the amount of nickel per catalytic unit re- quires knowledge of the amount of nickel, and the catalytic unit molecular weight. Studies incorporating the above in- formation have shown that the jack bean urease (19, 20) and the K. aerogenes enzyme (Chapter 2, 3; 90, 91) have 2 mol nickel/mol catalytic unit. For other bacterial ureases, 0.8—2.1 mol nickel/mol of large subunit has been calculated (reviewed by Mobley and Hausinger; 55) without determining the number of catalytic units per enzyme. Biophysical and spectrosc0pic analysis of the nickel center have only been done with the jack bean and K. aerogenes enzymes. UV- visible spectroscopy of K. aerogenes urease in the presence of thiol inhibitors is described in chapter 3 (91), and compared to the jack bean enzyme results (4, 18). X-ray absorption spectroscopy (XAS) of jack bean (1, 14) and K. aerogenes (Scott and Hausinger, in preparation) ureases have demonstrated that the nickel is coordinated by a mixture of nitrogen and oxygen ligands; however the precise identity of these ligands remains unknown. Lee et al. (47) observed in— creased reactivity of K. aerogenes apo—urease towards di— ethylpyrocarbonate at pH 6.5, consistent with (at least partial) histidine ligation of the nickel. Magnetic suscept- ibility measurements are consistent with at least a portion of the nickel being paramagnetic for both K. aerogenes (ap- pendix) and jack bean (17) ureases. In both cases, diamag— netism is observed in the presence of thiol compounds, in— dicating either that the the two nickel ions are antiferro— magnetically coupled, or that the individual nickel ions have become low spin (5:0) (17). Low temperature Magnetic Circular Dichroism (MCD) spectroscopic studies on K. aerogenes (Michael Johnson, unpublished experiments) and jack bean (25) ureases show that the two nickel become antiferromagnetically coupled in the presence of thiols, suggesting that these inhibitors bridge the two nickel at the active site. In summary, the active site of urease includes a novel, poorly characterized bi-nickel metallocenter. 0"! A ("1 f1 ’1 10 Kinetic parameters- The kinetic parameters of micro— bial ureases vary moderately and usually reflect the micro- organisms particular ecological niche. For example, the Kh values for urea are considerable higher for enzymes purified from organisms which invade the kidney (2-20 mM; 55) where the concentration of urea is high (0.4—0.5 M; 32), than from organisms which colonize the stomach lining (24, 54) where the concentration of serum urea is 1.5-3.0 mM (54). Urease Vkax is typically 1000-5000 umol urea hydrolyzed minflmg‘1 (reviewed by Mobley and Hausinger; 55). An important ex- ception is the urease of U. ureolyticum (V%ax ~ 90,000 umol min.“1 mg‘l; 82), which is at least 10-fold faster than any previously purified urease. The urease of this mycoplasma is thought to create an ammonia and H+ gradient, and the or- ganism may use this gradient to generate cellular energy (72). Most ureases have pH optima in the neutral to slightly basic range (55). Lactobacillus fermentum, however, has a urease with a pH optimum of 2 (41), which may reflect the ability of this organism to grow in acidic media. Inhibition of urease- The inhibition of urease has many potential applications. For example, derivatives of hydroxamic acid have been used as chemotherapeutic agents to reduce blood ammonia levels in patients with hyperammonemia (83), and to reduce urine ammonia and pH in patients with urinary infections (33, 62, 73). Serious side effects ll accompany the use of these compounds (27), and they have been suggested to be teratogenic and/or carcinogenic (73). Phosphoroamide derivatives (which are ~1000—fold more potent inhibitors) retard urinary stone growth (53, 85) and inhibit soil urease activity (16, 50, 61). In most instances where inhibition of urease is desired, the concentration of urea is present at levels which saturate the enzyme; therefore an ideal inhibitor would interact with urease by a non-com- petitive mechanism (57). Many determinations of the type of inhibition caused by these compounds reported non-competi- tive (26, 28, 29, 73) or mixed inhibition (28); other studies compared ISO values (the amount of inhibitor needed to obtain 50% inhibition of urease activity) without regard to the mechanism of inhibitor binding (27, 28, 34, 43). Acetohydroxamic acid and phenylphosphorodiamidate are slow- binding competitive inhibitors of K. aerogenes urease (Chapter 3, 91). To generate a useful comparison of the efficacy of these inhibitors using I50 values, one should also know the type of inhibition, the K5 and substrate concentration, and the enzyme concentration (11). Since these inhbiitors bind and dissociate from urease slowly, typical kinetic methods (derived assuming equilibrium or steady state conditions) simply reflect the amount of non- inhibited enzyme, thus appearing non-competitive (i.e. Vim, is decreased, K5 is not affected). The ionized form of hydroxamates are bidentate chela- tors of metals in solution and are used by some organisms as 12 siderophores (64). Ionized hydroxamates also inhibit metal- loproteins (ag.thermolysin, 65, 38) through bidentate bind— ing to the metallocenter. Changes in the UV-visible spectra of jack bean urease are consistent with their interaction with the active site nickel ion(s) (21). In addition, UV— visible spectra of one hydroxamic acid bound to jack bean urease, trans-cinnamoyl hydroxamate, was consistent with the ionized form of the hydroxamic acid being present at the active site, leading Zerner and colleagues to propose that hydroxamates bind to the urease nickel in a bidentate manner (21). In contrast, ionized hydroxamic acids do not interact with horseradish peroxidase; monodentate binding to the ac— tive site heme is thought to be the cause of hydroxamate in— hibition of this metalloenzyme (79). The binding of (neu- tral) acetohydroxamate to K. aerogenes urease is described in chapter 3 (91), and a distinct mechanism whereby the neu— tral hydroxamate bridges the two active site nickel (similar to the transition state of urea hydrolysis, Scheme 2) is proposed. Phosphoramides bind to the zinc in metalloproteases (such as carboxypeptidase A and thermolysin) and inhibit by acting as transition state analogues (42). Less is known about the interactions of phosphoramidates with urease. Zerner and colleagues, studying the reactivation of the jack bean enzyme inhibited with PPD, MBPT, and phosphoric tri- amide (2), observed similar reactivation rates with all 13 three compounds. From this observation, derivatives of phosphoroamide were proposed to be alternative substrates of jack bean urease; all three being hydrolyzed to a common in- hibitor, diamidophosphate. The slow-binding of phenylphos- phorodiamidate to K. aerogenes urease involved saturation kinetics similar to those seen for AHA (chapter 3; 91). PPD was also proposed to bind as a transition state analogue. Although no product analysis was done in either the jack bean or the bacterial urease study, slowly processed sub- strates are known to behave kinetically as slow—binding in— hibitors for several enzymes (57). In addition to the potent inhibition observed by hy— droxamates and phosphoramidates, many other compounds are known to inhibit urease by unknown mechanisms. In chapter 3 -(91), evidence is presented that indicates thiols, phos- phate, and boric and boronic acids are competitive inhibi— tors of K. aerogenes urease. Active site residues- Elucidation of an enzyme mechanism is not complete without examination of the active site amino acids which participate in substrate binding, catalytic turnover, or metal ligation. Chapters 2-5 include experiments designed to detect these amino acids, including effects of pH on kinetic parameters, inhibitor and inactiva- tor studies, and analysis of conserved sequences. Experiments using the well-studied plant urease demonstrated a slight increase in Kh upon deprotonation of an ionizable 14 group with pk; = 6.5 (22). Activity (Vhax) required depro- tonation of the group with pk; = 6.5 while retaining an enzyme group with pKh = 9.0 in a protonated state. Analo— gous studies have not been reported for microbial ureases, except those described in chapter 2 (91), which characterize the pH dependence of K. aerogenes urease: one enzyme group (pk; = 6.55) must be deprotonated, and a second enzyme group (pk; = 8.85) must be protonated for activity (90). The af- finity of enzyme for substrate is increased slightly below pH 6.5. While the identity of enzyme groups exhibiting these pkh's is not known, amino acid specific reagents have been used to probe their reactivitiy. Urease activity from plants (66, 71), algae (70) and bacteria (41, 63, 75, 92) is affected by thiol specific re- agents. Dixon et al. (22) proposed that jack bean urease has an essential thiol acting as a general acid during cata— lytic turnover (Scheme 2). This essential thiol, Cyssgz (87, 88), was reported to react with modifying reagents only in the ionized form.with a pKh = 9.0 (66). In contrast, the essential thiol in K. aerogenes urease interacts with some nearby amino acid, as yet unidentified, yielding a more com— plicated pH dependence of inactivation (Chapter 4; 92). The discrepancy between these two reports may lie in the methods used to determine the rate of essential thiol modification (see Chapter 4 for discussion); nevertheless, the analogous residue has been identified as the essential thiol in bac- terial urease: Cys319 of the large subunit (Chapter 5; 93). 15 Aside from an essential cysteine, the mechanism of urea hydrolysis as proposed by Dixon et al., (Scheme 2) invokes two additional active site amino acids: a carboxylic acid, and an unidentified base. Evidence supporting an essential carboxylic acid at the jack bean urease active site is based on unpublished experiments measuring urease activity in the presence of triethyloxonium ion. Indirect evidence of an anionic residue near the active site of bacterial urease comes from the comparison of various thiol inhibitor K} values (28, 34, 43, Chapter 3; 91) and comparison of the second order rate constants for several thiol modifying reagents (Chapter 4; 92). With regard to the general base, Sakaguchi et al. (77) have reported inactivation of jack bean urease by photo-oxidation of histidine residues using methylene blue; urease inhibitors prevented this inactiva- tion. Lee et al., (47) have seen loss of K. aerogenes urease activity using diethylpyrocarbonate, a modifying reagent which reacts with histidine residues. Comparison of published urease sequences from K. aerogenes (60), jack bean (88), P mirabilis (40), P. vulgaris (58), U. ureolyticum (7), and H. pyloris (46) demonstrate that ten histidine are conserved among these diversely represented species. One of these histidine resi- dues may correspond to the general base described above. In addition, several of these histidines may serve as ligands to the nickel, as described above. 16 The following chapters describe my studies on K. aerogenes urease. First I purify the nickel-containing enzyme 1070-fold and show that this and several other micro— bial ureases are heteropolymeric (Chapter 2; published in J. Biol. Chem., 262, 5963-5967). I then demonstrate that some competitive inhibitors directly interact with the nickel center, and propose models for the interaction of these inhibitors with the bi-metal active site (Chapter 3; published in J. Biol. Chem., 264, 15836—15842). Finally, I describe the unusual reactivity of the essential cysteine in urease (Chapter 4; published in J. Biol. Chem., 266, 10260- 10267), and I identify this amino acid as Cys319 in the a subunit (Chapter 5; 93). REFERENCES Alagna, L., Hasnain, S.S., Piggot, B., & Williams, D. J. (1984) The nickel ion environment in jack bean urease. Biochem. J. 220, 591-595 Andrews, R. K., Dexter, A., Blakeley, R. L., & Zerner, B. (1986) Jack bean urease (EC 3.5.1.5). VIII. On the inhibition of urease by amides and esters of phosphoric acid. J. Am. Chem. Soc. 108, 7124-7125 Beaton, J. D. (1978) Urea: its popularity grows as a dry source of nitrogen. Crops Soils 30, 11-14 Blakeley, R. L., Dixon, N. E., & Zerner, B. (1983) Jack bean urease VII. Light scattering and Ni (II) spectrum. Thiolate-nickel (II) charge-transfer peaks in the spectrum of the B-mercaptoethanol-urease complex. Biochim. et Biophys. Acta 774, 219-229 Blakeley, R. L., Hinds, J. A., Kunze, H. B., Webb, E. C., & Zerner, B. (1969) Jack bean urease (EC 3.5.1.5). Demonstration of a carbamoyl-transfer reaction and inhibition by hydroxamic acids. Biochem. 8. 1991—2000 Blakeley, R. L., & Zerner, B. (1984) Jack bean urease: The first nickel enzyme. J..Mol. Catalysis 23, 263-292 Blanchard, A. (1990) Ureaplasma urealyticum urease genes; use of a USA tryptophan codon. Mbl..Microbiol. 4. 669—676 Bremner, J. M., & Krogmeier, M. J. (1988) Elimination of the adverse effects of urea fertilizer on seed germination, seedling growth, and early plant growth in soil. Proc. Natl. Acad. Sci. 85, 4601-4604 17 10. 11. 12. 13. 14. 15. 16. 18 Byrnes, B. H., Savant, N. K., & Craswell, E. T. (1983) Effect of a urease inhibitor, phenylphosphorodiamide, on the efficience of urea applied to rice. Soil Sci. Soc. Am. J. 47, 270-274 Chen, X. G., Correa, P., Offerhaus, J., Rodriguez, E., Janney, P., Hoffmann, E., Fox, J., Hunter, F., & Diavolitsis, S. (1986) Ultrastructure of the gastric mucosa harboring campylobacter-like organisms. Am. J. of Clin. Path. 86, 575-582 Cheng, Y. C., & Prusoff, W. H. (1973) Relationship between the inhibition constant (Kfl and the concentration of inhibitor which causes 50 percent inhibition (Imfl of an enzymatic reaciton. Biochem. Pharmacol. 22, 3099-3108 Christians, 5., & Kaltwasser, H. (1986) Nickel- content of urease from Bacillus pasteurii. Arch. .Microbiol. 145, 51-55 Clark, P. A., & Wilcox, D. E. (1989) Magnetic properties of the nickel enzymes urease, nickel- substituted carboxypeptidase, A, and nickel-substituted carbonic anhydrase. Inorg. Chem. 28, 1326 Clark, P. A., Wilcox, D. E., & Scott, R. A. (1990) X— ray absorption spectroscopic evidence for binding of the competitive inhibitor 2-mercaptoethanol to the nickel sites of jack bean urease. A new Ni-Ni interaction in the inhibited enzyme. Inorg. Chem. 29, 579-581 Cooper, A. J., & Plum, F. (1987) Biochemistry and Physiology of Brain Ammonia. Phys. Rev. 67, 440-519. Creason, G. L., Schmitt, M. R., Douglass, E. A., & Hendrickson, L. L. (1990) Urease inhibitory activity associated with N—(n-butyl)thiophosphoric triamide is due to formation of its oxon analog. Soil Biol. Biochem. 22, 209-211 Ad. a s 17. 18. 19. 20. 21. 22. 23. 24. 19 Day, E. P., Peterson, J., Todd, M. J., & Hausinger, R. P. (1989) Saturation magnetization of the nickel centers in urease. J. Inorg. Biochem. 36, 305 Dixon, N. E., Blakeley, R. L., & Zerner, B. (1980a) Jack bean urease (EC 3.5.1.5). III. The involvement of active-site nickel ion in inhibition by B-mercapto- ethanol, phosphoramidate, and fluoride. Can. J. Biochem. 58, 481-488 Dixon, N. E., Gazzola, C., Blakeley, R. L., & Zerner, B. (1975a) Jack bean urease (EC 3.5.1.5). A metalloenzyme. A simple biological role for nickel? J. Am. Chem. Soc. 97, 4131-4133 Dixon, N. E., Gazzola, C., Watters, J. J., Blakeley, R. L., & Zerner, B. (1975b) Inhibition of jack bean urease (EC 3.5.1.5) by acetohydroxamic acid and by phosphoramidate. An equivalent weight for urease. J. Am. Chem. Soc. 97, 4130-4131 Dixon, N. E., Hinds, J. A., Fihelly, A. K., Gazzola, C., Winzor, D. J., Blakeley, R. L., & Zerner, B. (1980b) Jack bean urease (EC 3.5.1.5). IV. The molecular size and the mechanism of inhibition by hydroxamic acids. Spectrophotometric titration of enzymes with reversible inhibitors. Can. J. Biochem. 58. 1323-1334 Dixon, N. E., Riddles, P. W., Gazzola, C., Blakeley, R. L., & Zerner, B. (1980c) Jack bean urease (EC . 3.5.1.5). V. On the mechanism of action of urease on urea, formamide, acetamide, N-methylurea, and related compounds. Can. J. Biochem 58, 1335-1344 Dooley, C. P., & Cohen, H. (1988) The clinical significance of Campylobacter pylori. Annals of Int. .Med. 108, 70-79 Dunn, B. E., Campbell, G. P., Perez-Perez, G. I., & Blaser, M. J. (1990) Purification and character- ization of urease from Helicobacter pylori. J. Biol. Chem. 265, 9464-9469 25. 26. 27. 28. 29. 30. 31. 32. 33. 20 Finnegan, M. G., Kowel, A. T., Werth, M. T., Clark, P. A., Wilcox, D. E., & Johnson, M. K. (1991) Variable- temperature magnetic circular dichroism spectroscopy as a probe of the electronic and magnetic properties of nickel in jack bean urease. J. Am. Chem. Soc. 113, 4030-4032 Fishbein, W. N., & Carbone, P. P. (1965) Urease catalysis. II. Inhibition of the enzyme by hydroxyurea, hydroxylamine, and acetohydroxamic acid. J. Biol. Chem. 240, 2407-2414 Fishbein, W. N. (1982) Hydroxamic acids as urease inhibitors for medical and veterinary use. Chemistry and biology of hydroxamic acids (H. Kehl, ed), pp. 94- 103. Karger, Basel Gale, G. R. & Atkins, L. M. (1969) Inhibition of urease by hydroxamic acids. Arch. int. Pharmacodyn. 180. 289-298 Gatermann, S. John, J., & Marre, R. (1989) Staphylo- coccus saprophyticus urease: characterization and contribution to uropathogenicity in unobstructed urinary tract infection of rats. Infect. Immun. 57, 110-116 Goodwin, C. 8., Armstrong, J. A., Marshall, & B. J. (1986) Campylobacter pyloridis, gastritis, and peptic ulceration. J. Clin. Path. 39, 353-365 Graham, D. Y., & Klein, P, D. (1987) Campylobacter pyloridis gastritis: the past, the present, and speculations about the future. Am. J. Gastroenter. 82. 283-286 Griffith, D. P., Musher, D. M., & Hin, C. (1976) Urease. The primary cause of infection-induced urinary stones. Invest. urol. 13, 346-350 Griffith, D. P. (1978) Struvite stones. Kidney int. 13' 372‘82 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 21 Hase, J. & Kobashi, K. (1967) Inhibition of Proteus vulgaris urease by hydroxamic acids. J. Biochem. 62, 293-299 Hausinger, R. P. (1987) Nickel utilization by microorganisms. Microbiol. Rev. 51, 22-42 Hausinger, R. P. (1990) Mechanisms of metal ion incorporation into metalloproteins. Biofactors 2, 179-184 Hazell, S. L., & Lee, A. (1986) Campylobacter pyloridis, urease, hydrogen ion back diffusion, and gastric ulcers. Lancet ii 15-17 Holmes, M. A., & Matthews, B. W. (1981) Binding of hydroxamic acid inhibitors to crystalline thermolysin suggests a pentacoordinate zinc intermediate in catalysis. Biochem. 20, 6912-6920 Hu, L., Nicholson, E. B., Jones, B. D., Lynch, M. J., & Mobley, H. L. T. (1990) .Mbrganella morganii urease: purification, characteization, and isolation of gene (sequences. J. Bacteriol. 172, 3073-3080 Jones, E. D. & Mobley, H. L. T. (1988) Proteus mirabilis urease: nucleotide sequence determination and comparison with jack bean urease. J. Bacteriol. 171. 6414-6422 Kakimoto, S., Sumino, Y., Kawahara, K., Yamazaki, E., & Nakatsui, I. (1990) Purification and characterization of acid urease from Lactobacillus fermentum. Appl. .Microbiol. Biotechnol. 32, 538-543 Kam, C. M., Nishino, N., & Powers, J. C. (1979) Inhibition of thermolysin and carboxypeptidase A by phosphoramidates. Biochem. 18. 3032-3038 Kobashi, K., Kumaki, K., & Hase, J. (1971) Effect of acyl residues of hydroxamic acids on urease inhibition. Biochim. Biophys. Acta. 227, 429-441 44. 45. 46. 47. 48. 49. 50. 51. 52. 22 Kobashi, K., Munakata, K., Takebe, S., & Hase, J. (1980) Therapy for urolithiasis by hydroxamic acids. II. Urease inhibitory potency and urinary excretion rate of hippurohydroxamic acid derivatives. J. Pharm. Dyn. 3. 444-450 Krogmeier, M. J., McCarty, G. W., & Bremner, J. M. (1989) Potential phytotoxicity associated with the use of soil urease inhibitors. Proc. Natl. Acad. Sci. 86, 1110-1112 Labigne, A., Cussac, V., & Courcous, P. (1991) Shuttle cloning and nucleotide sequences of Helicobacter pylori genes responsible for urease activity. J. Bacteriol 173, 1920-1931 Lee, M. Y., Mulrooney, S. B., & Hausinger, R. P. (1989) Purification, characterization, and in-vivo reconstitution of Klebsiella aerogenes urease apoenzyme. J. Bacteriol. 172, 4427-4431 Mak, A. L., & Trevan, M. D. (1988) Urea as a nitrogen source for calcium-alginate immobilized Chlorella. Enzyme Microb. Tbchnol. 10, 207—213 Mak, A. L., Trevan, M. D., Rogers, L. J., & Gallon, J. R. (1989) Urea as a nitrogen source for immobilized Chlorella. Biochemistry of the algae and cyanobac- teria. Proc. Phytochem. Soc. Eur. 1989, 362-363 Martens, D. A., & Bremner, J. M. (1984) Effectiveness of phosphoroamides for retardation of urea hydrolysis in soils. Soil. Sci. Soc. Am. J. 48, 302-305 McLean, R. J. C., Nickel, J. C., Beveridge, T. J., & Costerton, J. W. (1989) Observations of the ultra- structure of infected kidney stones. J..Med. Microbiol. 29, 1-7 McLean, R. J. C., Nickel, J. C., Cheng, K. J., & Costerton, J. W. (1988) The ecology and pathogenicity of urease producing bacteria in the urinary tract. CRC Crit. Rev..Microbiol. 16, 37-79 r!» .4v wa .s' 744 .s. 53. 54. 55. 56. 57. 58. 59. 60. 61. 23 Millner, O. E., Andersen, J. A., Appler, M. E., Benjamin, C. E., Edwards, J. G., Humphrey, D. T., & Shearer, E. M. (1982) Flurofamide: a potent inhibitor of bacterial urease with potential clinical utility in the treatment of infection induced urinary stones. J. urol. 127, 346—350 Mobley, H. L. T., Cortesia, M. J., Rosenthal, L. E., & Jones, B. D. (1988) Characterization of urease from Campylabacter pylori. J. Clin. Microbiol. 26, 831-836 Mobley, H. L. T., & Hausinger, R. P. (1989) Microbial ureases: significance, regulation, and molecular characterization. .Microbial. Rev. 53, 85-108 Mobley, H. L. T., & Warren, J. W. (1987) Urease- positive bacteriuria and obstruction of long-term urinary catheters. J. Clin. Microbiol. 25, 2216-2217 Morrison, J. F. & Walsh, C. T. (1988) The behavior and significance of slow-binding enzyme inhibitors. Adv. Enzymal. 61, 201-301 Morsdorf, G. & Kaltwasser, H. (1990) Cloning of the genes encoding urease from Proteus vulgaris and sequencing of the structural genes. FEMS Microbiol. Lett. 66, 67-74 Mulrooney, S. B., Lynch, M. J., Mobley, H. L. T., & Hausinger, R. P. (1988) Purification, characterization, and genetic organization of recombinant Providencia stuartii urease expressed in Eschericia cali. J. Bacterial. 170, 2202-2207 Mulrooney, S. B., & Hausinger, R. P. (1990) Sequence of the Klebsiella aerogenes urease genes and evidence for accessory proteins facilitating nickel incorporation. J. Bacterial. 172, 5837-5843 Mulvaney, R. L., & Bremner, J. M. (1981) Control of urea transformations in soils. Sail Biochemistry, vol. 5. (E. A. Pau and J. N. Ladd, eds.) pp. 153-196, Marcel Dekker, Inc., New York 62. 63. 64. 65. 66. 67. 68. 69. 70. 24 Munakata, K., Kobashi, K., Takebe, S., & Hase, J. (1980) Therapy for urolithiasis by hydroxamic acids. III. urease inhibitory potency and urinary excretion rate of N-acylglycinohydroxamic acids. J. Pharm. Dyn. 3. 451-456 Nakano, H., Takenishi, S., & Watanabe, Y. (1984) Purification and properties of urease from Brevibacterium ammoniagenes. Agric. Biol. Chem. 48, 1495-1502 Nielands, J. B. (1967) Hydroxamic acids in Nature. Science 156, 1443-1447 Nishino, N. & Powers, J. C. (1979) Design of potent reversible inhibitors for thermolysin. Peptides containing zinc coordinating ligands and their use in affinity chromatography. Biochem. 18, 4340-4347 Norris, R., & Brocklehurst, K. (1976) A convenient method of preparation of high-activity urease from Canavalia ensifarmis by covalent chromatography and an investigation of its thiol groups with 2,2'-dipyridyl disulphide as a thiol titrant and reactivity probe. Biochem. J. 159. 245-257 Piskin, E., & Chang, T. M. S. (1981) Urea and ammonia removal from dialysate in an oxygenator: effects of dialysate and air flow rates. J..Membrane Sci. 9, 343-350 Polacco, J. C., & Havir, E. A. (1979) Comparisons of soybean urease isolated from seed and tissue culture. J. Biol. Chem 254 1707-1715 Pugh, K. B., & Waid, J. S. (1969) The influence of hydroxamates on ammonia loss from various soils treated with urea. Soil Biol. Biochem. 1, 207-217 Rai, A. K., & Singh, S. (1987) Urease of blue-green algae (cyanobacteria) Anabene dalialum and Anacystis nidulans. Curr. Microbiol. 16, 113-117 4. u p N. Nu -. k9 71. 72. 73. 74. 75. 76. 77. 78. 79. 25 Riddles, P. W., Andrews, R. K., Blakeley, R. L., & Zerner, B. (1983) Jack bean urease. VI. Determination of thiol and disulfide content. Reversible inactivation of the enzyme by the blocking of the unique cysteine residue. Biochim. Biophys. Acta. 743. 115-120 Romano, N., Tolone, G., Ajello, F., & LaLicata, R. (1980) Adenosine 5'-triphosphate synthesis induced by urea hydrolysis in Ureaplasma urealyticum. J. Bacteriol. 144, 830-832 Rosenstein, I. J. M., & Hamilton-Miller, J. M. T. (1984) Inhibitors of urease as chemotherapeutic agents. CRC Crit. Rev. Microbial. 11, 1-12 Rubin, R. H., Tolkoff—Rubin, N. E., & Cotran, R. S. (1986) Urinary tract infection, pyelonephritis, and reflux nephropathy. The Kidney. (B. M. Bremner & F. C. Rector, eds) pp. 1085-1141. W. B. Saunders Co., Philadelphia Saada, A. B., & Kahane, I. (1988) Purification and characterization of urease from ureaplasma urealyticum. Zbl. Bakt. Hyg. A 269, 160-167 Sabbaj, J., Sutter, V. L., & Finegold, S. M. (1970) Urease and deaminase activities of fecal bacteria in hepatic coma. Antimicrob. Agents Chemotherapy. 1970, 181-185 Sakaguchi, K., Mitsui, K., Kobashi, K., & Hase, J. (1983) Photo-oxidation of jack bean urease in the presence of methylene blue. J. Biochem. 93, 681-686 Samtoy, B. & DeBeukelar, M. M. (1980) Ammonia encephalopathy secondary to urinary tract infection with Proteus mirabilis. Pediatrics 65. 294-297 Schonbaum, G. R. (1973) New complexes of peroxidases with hydroxamic acids, hydrazides, and amides. J. Biol. Chem. 248, 502-511 .IV «(A u shy 80. 81. 82. 83. 84. 85. 86. 87. 88. 26 Slomiany, B. L., Bilski, J., & Murty, U. L. (1987) Campylabacter pyloridis degrades mucin and undermines gastric mucosal integrity. Gastraenteralogy 92, 1645 Smoot, D. T., Mobley, H. L. T., Chippendale, G. R., Lewison, J. F., & Resau, F. H. (1990) Helicobacter pylori urease activity is toxic to human gastric epithelial cells. Infec. and Immun. 58, 1992-1994 Stemke, G. W., Robertson, J. A., & Nhan, M. (1987) Purification of urease from Ureaplasma urealyticum. Can. J..Micrabiol. 33, 857-862 Summerskill, J., Thorsell, E., Feinberg, J. H., & Aldrete, J. S. (1967) Effects of urease inhibition in hyperammonemia: clinical and experimental studies with acetohydroxamic acid. Gastraent. 54, 20-26 Sumner, J. B. (1926) The isolation and crystallization of the enzyme urease. J. Biol. Chem. 69. 435-441 Takebe, S., Numata, A., & Kobashi, K. (1984) Stone formation by Ureaplasma urealyticum in human urine and its prevention by urease inhibitors. J. Clin. .Microbiol. 20, 869-873 Takeuchi, H., Takayama, H., Konishi, T., & Tomoyoshi, T. (1984) Scanning electron microscopy detects bacteria within infection stones. J. Urol. 132, 67-69 Takishima, K., Mamiya, G., & Hata, M. (1983) Amino acid sequence of a peptide containing an essential cysteine residue of jack bean urease. Frontiers in biochemical and biophysical studies of proteins and membranes (T. Y. Liu, S. Sakakibara, A. N. Schecter, K. Yagi, H. Yajima, and K. T. Yasunobu, eds) Elsevier, N.Y. Takishima, K., Suga, T., & Mamiya, G. (1988) The structure of jack bean urease. The complete amino acid sequence, limited proteolysis and reactive cysteine residues. Eur. J. Biochem. 175, 151-165 89. 90. 91. 92. 93. 94. 27 Thirkell, D., Myles, A. D., Precious, B. L., Frost, J. S., Woodall, J. C., Burdon, M. G., & Russell, W. C. (1989) The urease of Ureaplasma urealyticum. J. Gen. Microbiol. 135. 315-323 Todd, M. J., & Hausinger, R. P. (1987) Purification and characterization of the nickel-containing multicomponent urease from Klebsiella aerogenes. J. Biol. Chem. 262, 5963-5967 Todd, M. J., & Hausinger, R. P. (1989) Competitive inhibitors of Klebsiella aerogenes urease. J. Biol. Chem. 264. 15835-15842 Todd, M. J., & Hausinger, R. P. (1991a) Reactivitiy of the essential thiol of Klebsiella aerogenes urease. Effect of pH and ligands on thiol modification. J. Biol. Chem. 266, 10260-10267 Todd, M. J., & Hausinger, R. P. (1991b) Identification of the essential cysteine residue in Klebsiella aerogenes urease. J. Biol. Chem. submitted Winkler, R. G., Blevins, D. G., Polacco, J. C., & Randall, D. D. (1988) Ureide catabolism in nitrogen fixing legumes. Trends Biochem. Sci. 13, 97-100 PURIFICATION AND CHARACTERIZATION OF THE NICKEL-CONTAINING HULTICOKPONENT UREASE FROM KLEBSIELLA AEROGENES ABSTRACT Klebsiella aerogenes urease was purified 1,070-fold with a 25% yield by a simple procedure involving DEAE- Sepharose, phenyl-Sepharose, Mono Q, and Superose 6 chrom— atographies. The enzyme preparation was comprised of three polypeptides with estimated M1. = 72,000, 11,000, and 9,000 in a “25474 quaternary structure. The three components re- mained associated during native gel electrophoresis, Mono Q chromatography, and Superose 6 chromatography despite the presence of thiols, glycols, detergents, and varied buffer conditions. The apparent compositional complexity of K. aerogenes urease contrasts with the simple well-character- ized homohexameric structure for jack bean urease (Dixon, N. E., Hinds, J. A., Fihelly, A. K., Gazzola, C., Winzor, D. J., Blakeley, R. L., and Zerner, B. (1980) Can. J. Biochem. 58, 1323-1334); however, heteromeric subunit compositions were also observed for the enzymes from Proteus mirabilis, Sparosarcina ureae, and Selenamanas ruminantium. K. aerogenes urease exhibited a K5 for urea of 2.8 i 0.6 mM and a Vhaxiof 2,800 i 200 umol of urea min’l mg‘l at 37'C in 25 mM N-2-hydroxyethylpiperazine-N'-2-ethanesulfonic acid, 28 29 5.0 mM EDTA buffer, pH 7.75. The enzyme activity was stable in 1% sodium dodecyl sulfate, 5% Triton X-100, 1 M KCl, and over a pH range from 5 to 10.5, with maximum activity ob- served at pH 7.75. Two active site groups were defined by their pKQ values of 6.55 and 8.85. The amino acid composi- tion of K. aerogenes urease more closely resembled that for the enzyme from Brevibacter ammoniagenes (Nakano, H., Takenishi, S., and Watanabe, Y. (1984) Agric. Biol. Chem 48, 1495-1502 than those for plant ureases. Atomic absorp— tion analysis was used to establish the presence of 2.1 i 0.3 mol nickel/mol 72,000-dalton subunit in K. aerogenes urease . ‘ ! VPL «an s\k Introduction Urease is a nickel-containing enzyme which catalyzes the hydrolysis of urea to form carbon dioxide and ammonia (1,2). The archetype urease, isolated from jack bean, has been intensively studied since 1926 when Sumner first crystallized this protein (3) . Jack bean urease (M1. = 590,000) is a hexamer of identical subunits (4) of known amino acid sequence (AL.= 90,790) (5). This plant enzyme contains two nickel ions per subunit (6), and the metal ions are apparently coordinated by oxygen and nitrogen ligands (7). Whereas soybean urease appears to be analogous in size, structure, and nickel content (8), microbial ureases are distinct from the jack bean enzyme. For example, ureases from Brevibacter ammoniagenes and Bacillus pasteurii are smaller in subunit size (Afi.= 67,000 and 65,000) and native size (Afi.= 200,000 and 230,000) and possess a single nickel ion per subunit (9,10). The Selenamanas ruminantium urease is also smaller (native hfi.= 360,000), but like the plant enzymes it contains two nickel ions per subunit (hfi.= 70,000) (11). Active site differences may also exist in the microbial and plant enzymes as shown by their differences in susceptibility to various inhibitors (12). Herein we extend the comparison of plant and microbial ureases by characterizing the isolated Klebsiella aerogenes 3O 31 enzyme. Previously, Friedrich and Magasanik (13) purified K. aerogenes urease 24-fold and examined the cellular re- gulation of this enzyme. In addition, Kamel and Hamed (14) described several properties of 150-fold-purified urease from the related organism, Aerabacter aerogenes PRL—R3. In this report, the nickel-containing K. aerogenes urease is intensively characterized and shown to have a novel urease structure comprised of three distinct polypeptide chains. EXPERIMENTAL PROCEDURES Bacterial growth conditions- Klebsiella aerogenes CG253, obtained from Boris Magasanik and Alexander Ninfa (Massachusetts Institute of Technology), was cultured at 37’C in minimal media which contained 10 g sucrose, 10.5 g K¢HPO,, 4.5 g KHZPO4, 0.2 g MgSOz-7Hy3, 2.5 ml trace min- erals, and 0.24 g urea (filter sterilized) per liter. The trace mineral solution contained, per 100 ml, 20 mg ZnSO4o7HéO, 10 mg H3BO3, 10 mg NaSeO3, 10 mg NaMoO4-2H¢O, 10 mg NiC12-6H20, 5 mg CuSO4-5HZO, 2 mg AlK(SO4)2-12HZO and 1 ml H2804 (15). Large scale aerobic batch culture was performed by using a New Brunswick 25L Microferm fermenter with rapid agitation. The culture was grown to an absorbance (600 nm) of 1.8 and harvested by using a Pellicon Cassette system concentrator (Millipore). The cells (80 to 108 g wet weight per 22 L) were washed with 20 mM potassium phosphate, 1 mM 32 EDTA, 1 mM B-mercaptoethanol, pH 7.0 (abbreviated PEB) buffer and either used immediately or stored at -20’C. Assays- Urease was assayed by measuring the rate of release of ammonia from urea. The released ammonia was converted to indophenol whose absorbance was monitored at 625 nm (16). Specific activity of urease was defined as umol urea hydrolyzed min‘lmg“l at 37'C in 25 mM HEPES (N-2- hydroxyethylpiperazine-N'-2-ethane sulfonic acid), 5.1 mM EDTA, pH 7.75 buffer containing 25 mM urea. Protein was assayed as described by Lowry et al. (17) with bovine serum albumin as the standard. In addition, the absorbance at 280 nm was used to quantitate highly purified samples of urease. An extinction coefficient of 8.5 at 280 nm was obtained for a 1% solution of the isolated enzyme, based on the protein concentration obtained by using the Lowry method. All UV/visible absorbance determinations were made by using a Gilford Response spectrophotometer. Preparation of crude extract- The cells were suspended in an equal volume of PEB buffer containing 1.0 mM toluene sulfonyl fluoride and disrupted by two passages through a French pressure cell at 16,000 psi. In some cases, other protease inhibitors were used including bestatin, pep- statin, leupeptin, aprotinin, and phosphoramidone. Mem- branes and cellular debris were removed by centrifugation at 10,000 x g for 60 min at 4‘C. All isolation procedures, 33 except that using the Fast Protein Liquid Chromatography (FPLC) system, were carried out at 4°C and using PEB buffer with the indicated additions. SDS-palyacrylamide gel electrophoresis- The molecular weight for the urease polypeptides and the purity of samples were assessed by using SDS-polyacrylamide gel electro- phoresis in a Hoefer SE 600 electrophoresis apparatus as described by Laemmli (18). The samples were denatured for 5 min at 100'C in 0.0625 M Tris buffer, pH 6.8 containing 3% SDS, 10% glycerol, and 5% B-mercaptoethanol. These samples were then electrophoresed at 30 mA constant current through a 3% or 4.5% acrylamide stacking gel and a 10%, 12% or 20% acrylamide running gel, each 1.5 mm in thickness. To im- prove molecular weight estimates for small polypeptides, 20% acrylamide gels possessed a 200:1 ratio of acrylamide to methylene bisacrylamide. In some cases, 5% to 15% acryl- amide gradient gels were prepared to enhance the resolution of the entire range of peptides. Standards (Bio-Rad, Cal- biochem, Sigma) used for comparison were: myosin, hfi.= 200,000; B-galactosidase, Mr = 116,250; phosphorylase b, Mr = 92,500; jack bean urease, Afi.= 90,790 (5); bovine serum albumin, Mr = 66,200; ovalbumin, Mr 2 45,000; carbonic anhydrase, Mr = 31,000; chymotrypsinogen, Mr = 25,000; soybean trypsin inhibitor, Mr = 21,500; myoglobin, Mr = 16,950; a-lactalbumin, Mr = 14,200; lysozyme, M1. = 14,000; cytochrome C, hfi.= 12,500; and myoglobin cyanogen bromide 34 fragments, Nfi.= 14,000, 8,160, 6,210, and 2,510. Gels were stained with Coomassie Brilliant blue or Silver stain (19) to detect proteins and with dansyl hydrazine to detect glycoproteins (20). Native gel electrophoresis— Nondenaturing gel electro— phoresis was carried out by using the buffers described by Laemmli (18) without SDS. Samples were electrophoresed at 35 mA constant current into the 3% acrylamide stacking gel and 5%, 7%, or 9% acrylamide running gel, 1.5 mm in thick- ness. Urease was detected in the gels either by a phenol red activity stain analogous to the cresol green stain described by Blattler et al. (21) or by the nitroblue tetrazolium activity stain described by Fishbein (22). Other proteins were detected by Coomassie brilliant blue or Silver stain methods. Native molecular weight determination- The native molecular weight of K. aerogenes urease was estimated by two methods. Native gel electrophoresis was carried out on samples and standards by using several gels of varied acryl- amide concentration (5% to 9%). The method of Zwaan was used to estimate the molecular weight by plotting the native protein size as a function of the ratio of percent migration versus tracking dye in the two gels (23). The standards used in the gel method included carbonic anhydrase, hfi.= 31,000; ovalbumin, Mr = 45,000; ovalbumin dimer, Mr = 35 90,000; bovine serum albumin (BSA) monomer, NQ.= 66,200; BSA dimer, M = 132,000; BSA trimer, Mr = 198,600; myosin, I." Mr = 200,000; and B-galactosidase, Mr = 465,000. In addition, the molecular weight for native urease was deduced by using Superose 6 (1.0 x 30 cm, Pharmacia) gel filtration chromatography. Standards included thyroglob- ulin, Mr = 670,000; ferritin, Mr = 470,000; y-globulin, Mr = 158,000; ovalbumin, Mr = 44,000; myoglobin, Mr = 17,000; and vitamin B-12, AL.==14350. Samples were eluted at 0.5 ml min'1 in PEB buffer containing 0.1 M KCl while monitoring the absorbance at 280 nm. Stability studies- Urease (7 ug mail) was incubated at O-2'C in a 10-fold dilution of assay buffer containing var- ious concentrations of detergents, glycols, thiols, salts, or mixtures of these substances. Aliquots from these incubations were diluted 200-fold into the routine assay buffer and assayed for activity. (pH Studies- The enzyme stability was assessed at pH values from 4 to 11 by incubating urease in buffers at various pH values for 30 min then assaying aliquots of the mixture in a 40-fold volume of the routine assay buffer containing 2.5 mM to 50 mM urea. The test buffers included acetate (pH 4-6), MES (pH 5-7), CHES (pH 8.5-10.5), or CAPS (pH 9-12) at a concentration of 50 mM. In addition, all buffers contained 10 mM EDTA. 36 The effects of pH on Km (urea) and Vmax were estab- lished by assaying urease activity in buffers containing 25 mM buffer, 5 mM EDTA and 2.5 mM to 50 mM urea, at the in- dicated pH values. The rate of urea hydrolysis was linear at each pH value at which the enzyme was stable. Buffers included the four mentioned above, HEPES (pH 6.5-9.0), and phosphate (pH 5.0-11.0). Amino acid analysis- Protein samples were hydrolyzed in vacuo in 6 N HCl at 110°C for 24, 48, and 72 hr. Dup- licate samples were analyzed by using a Beckman model 119CL automatic amino acid analyzer with ninhydrin detection system. The instrument was equipped with a model 126 data system for peak integration at 440 and 570 nm. Cysteine was quantitated as carboxymethyl cysteine for samples which were denatured, reduced and alkylated by using iodoacetic acid (2%). Threonine and serine values were extrapolated to zero time of hydrolysis. Tryptophan was quantitated spectro- photometrically by the absorbance at 280 nm; the extinction coefficient was taken to be 5700 M‘hxwl for Trp and 1300 M‘lcm‘l for Tyr. Ni analysis- The nickel content of samples was quanti- tated by using a Perkin-Elmer PE-500 atomic absorption spec- trophotometer equipped with an HGA 500 graphite furnace and an AS-l autosampler. Samples were hydrolyzed in 1 M nitric I r-\ a: u." ‘o a \V 3AM 37 acid, evaporated, and resuspended in 50 mM HNO3. Nickel standards, which in some cases contained bovine serum albumin to provide a protein matrix, were treated iden- tically to the urease samples. Aliquots of each sample (20 ul) were dried at 120'C, charred at 1200°C, and atomized at 2700'C. Nickel analysis was carried out using the in- strument's background correction mode, and either peak height or peak area were monitored. RESULTS Urease purification- Extracts from 190 g of cell (wet weight) were chromatographed on a column (2.5 x 15 cm) of DEAE-Sepharose (Pharmacia) as shown in Figure 1. A 400 ml linear gradient from 0 M to 1 M KCl in PEB buffer was used to recover urease as a single peak of activity at approx- imately 0.45 M KCl. Fractions containing peak urease activity were adjusted to 1 M KCl and chromatographed on a column (2.5 X 15 cm) of phenyl-Sepharose (Pharmacia) by using step elution. The column was washed with 150 ml of 1 M KCl in PEB and activity was eluted with 100 ml PEB buffer. As illustrated in Figure 2, a broad peak of urease activity was obtained, well separated from the major peak of protein. The peak urease fractions were pooled, diluted with an equal volume of PEB buffer and chromatographed on a Mono Q column 38 = T «000 ' 7'. A ‘: . E -‘ 1.0 O o . I.» 1' o ' ,u (600 s u ‘0 : '. ' E ‘I' . I - C . ' 9 O . . 4 g . 5 i v - C o 1 ‘00 I z ‘ . I g - 3 : I > 10 5 ‘ o r" :- '=‘ '3'” g u U ..... ' u 4 5!. 2° ‘ J 200 1 O. ‘ 6.. ‘ ..... . Ln" 4 , —‘ 0° 360 400 500 500 7oo .00 VOLUME (ml) Figure 1. DBAB-Sepharoae chromatography of x. aerogenes urease. Cell extracts were chromatographed on a column of DEAE-Sepharose by using a linear KCl gradient as indicated ( ' '° ° ). Fractions (5.5 ml) were monitored for absorbance at 280 nm after diluting aliquots 20-fold (O————O), and samples were assayed for urease activity (O-— - »O). The pooled fractions are indicated by the bar. 39 |.O M KCI PEB IOOO O O 800 N 8 600 C O a b o 3 400 4 200 VOLUME ( ml ) Figure 2. Phanyl-Sepharosa chromatography of K.. aerogenes urease. The pooled fractions from Figure 1 were adjusted to 1 M KCl and chromatographed on a column of phenyl-Sepharose by using step elution, decreasing the concentration of KCl as indicated while monitoring the absorbance at 280 nm (0 ———-O). Samples (3 ml) were assayed for urease activity (0 - - - O) and the pooled fractions are indicated by the bar. ACTIVITY (umol min" ml") NH). CEF-h<~ 40 (1.0 X 10 cm, Pharmacia) As shown in Figure 3, a KCl gradient was used to resolve urease from several contam- inants. The Mono Q fractions containing the highest urease activity were concentrated to 0.2 ml by using a Centricon-lO (Amicon) and chromatographed on a column (1.0 X 30 cm) of Superose 6 (Pharmacia) as illustrated in Figure 4. This final purification step served to desalt the enzyme and to provide an estimate of the native molecular weight (vide infra). The purification of K. aerogenes urease is summarized in Table 1. Table 1 Urease purification from K. aerogenes 'Rmfl Emnmm Pun’fication step Specific Activity Purification Total Activity Protein Recovery (mmole urea min'1mg") (told) (mmol min") (mg) (96) Cell extracts 2.06 1 19500 9460 100 DEAE-Sepharose 81.1 39 11800 145 60.5 PhenyI-Sepharose 574 279 5420 9.45 27.8 Mono Q 1170 568 5080 4 34 26.1 Superose 6 2200 1070 4770 217 24.5 41 8“. .‘b - U ”h I . 7 n “In ' .D’ ' ' “I 9 “b - .I. . 3 08)- a - p ”1- 2 «.80. .0)- Figure 3. none Q chromatography of.x. aerogenes urease. The pooled fraction from Figure 2 was subjected to Mono Q FPLC by using a KCl gradient as indicated (' °' ). The effluent was monitored for absorbance at 280 nm (————7 and the fractions were assayed for urease activity (O - - - O). The pooled fractions are indicated by the bar. 42 C22] " f .. 1.54 I: .6000 g o 00 m '1 r N - E 1v j E 3 1.0L 24000 g 3 i: a 0 °- 1. m 32 3, 1' : ‘3 0.5L : f -2000 3.” 3 °. .5. I - h o ' U j L‘A ( E=§= 1 . L . . it 0| 611 G 20 N u u 0 VOLUME (m0 Figure 4. Superose 6 chromatography of x; aerogenes urease. The urease containing pool from Figure 3 was con- centrated to 0.2 ml and chromatographed on a column of Su- perose 6 resin. The effluent was monitored for absorbance at 280 nm (———4 and the fractions were assayed for activity (O---O). 43 SDS-polyacrylamide gel electrophoresis of K. aerogenes urease- K. aerogenes urease was found to contain three distinct polypeptides (Afl.= 72,000 1 2,000; 11,000 r 2,000; and 9,000 r 2,000) by using SDS-polyacrylamide gel electro- phoresis (Figure 5). Gels comprised of less than 10% acryl- amide did not resolve the smaller components from the bromo— phenol blue dye front; thus 12% or 20% acrylamide gels were used to estimate the molecular weight of the small compon- ents. The integrated intensities of the three bands from Coomassie blue stained gels were used to compare the compon- ent ratios. This method is suspect because individual polypeptides may have different affinities for the dye; nevertheless, this procedure yielded ratios for the large:small:smallest bands of 1:2:2. None of the components were glycopeptides, as demonstrated by a failure to stain using dansyl hydrazine, a carbohydrate specific stain. Native molecular weight for K. aerogenes urease- By using the method of Zwaan (23) to analyze mobilities on non- denaturing gels, the native molecular weight for K. aerogenes urease was estimated as 190,000 r 20,000 daltons. This value was not significantly altered if the enzyme was incubated at room temperature in 1.0 M KCl, 5% SDS, 5% Triton X-100, 50% ethylene glycol, 50% glycerol, 50% sucrose, 5% B—mercaptoethanol, 20% dimethylsulfoxide, or mixtures of these substances prior to electrophoresis. Superose 6 gel filtration chromatography was used as a 44 ZCHD III! ||6 . ~ F 92.5 ‘- 45 "" I... 31 --s -' ‘3 21.5 - 14.4 -- “1...... ______,_ kD Std Kn. Rm. Su Sr Figure 5. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of several microbial ureases. Standard proteins (2 #9) and ureases from K. aerogenes (K. a.), P. mirabilis (P. m.), S. ureae (S. u.), and S. ruminantium (S. r.) (8 pg each) were denatured and electrophoresed on a 5- 15% gradient gel, and then stained with Coomassie Brilliant Blue. Details of these methods are described under "Experimental Procedures". 45 second method to estimate the native molecular weight for urease. Assuming that the ratio of molecular weight to Stoke's radius is similar for the enzyme and the standards, the K. aerogenes urease molecular weight is 260,000 r 50,000 daltons. Attempts to resolve the three urease components- The presence of three polypeptide components in K. aerogenes urease is unprecedented when compared with previously pub- lished urease preparations. To examine this discrepancy, intensive efforts were carried out to further resolve the enzyme by electrophoretic and chromatographic techniques. Two-dimensional gel electrophoresis, in which the urease band from native gel electrophoresis was subsequently de- natured and run on an SDS polyacrylamide gel, was used to demonstrate that all three polypeptides remain associated during native gel electrophoresis. Furthermore, prior in- cubation of urease in the presence of 1 M salt, 5% SDS, 5% Triton X-100, 50% ethylene glycol, 50% glycerol, 50% sucrose, 5% B-mercaptoethanol, 20% dimethylsulfoxide, or mixtures of these substances had no apparent effect on component association as demonstrated by native gel electro- phoresis. Ureolytic fractions obtained by using Superose 6 chromatography equilibrated with PEB buffers containing 0.5 M KCl, 0.1% SDS, or 3% B-mercaptoethanol consistently pos- sessed all three components upon SDS-polyacrylamide gel electrophoretic analysis. Mono Q chromatography of active 46 urease under a variety of buffer conditions (pH values from 6.0 to 9.0, the use of 20 mM B-mercaptoethanol, and elution with 0.5% triton X-100 in the buffers) also was incapable of resolving the three urease components. Thus, the three polypeptides remained associated under all conditions at which the enzyme activity was stable. Resolution of the large polypeptide has only been achieved by Superose 6 gel filtration chromatography in the presence of 6 M guanidine- HCl; however, enzyme activity was lost under these condi- tions. Enzyme stability studies- Attempts to resolve the three urease components included the use of rather harsh conditions; yet, the enzyme retained activity in most of these experiments. The stability of K. aerogenes urease was quantitatively examined under varied buffer conditions at 0- 2°C, as summarized in Table 2. The enzyme was found to retain nearly full activity for over 200 h despite the presence of salts, glycols, thiols, and detergents. Glycols, and perhaps 1 M KCl, enhanced the ureolytic activity at initial times; however 20-30% of the initial values were lost after 200 h, resulting in little net change from the control. The enzyme activity appeared to be more stable in Triton X-100 than in SDS; nevertheless, the enzyme could tolerate high concentrations of both detergents. In- cubations containing B-mercaptoethanol exhibited an initial drop in activity, consistent with competitive inhibition by 47 Table 2 Stability of K. aerogenes urease Buffer additionsz Percent activity remaining1 0 h 24 h 200 h None 100 106 111 1 M KCl 119 100 97 1 M KCl, 20% glycerol 131 107 91 20% glycerol 133 115 104 20% ethylene glycol 131 119 106 200 mM B-ME3 92 104 97 l % SDS 95 118 61 1 % SDS, 200 mM B—ME 89 99 95 5 % Triton X-100 85 112 102 5 % Triton X-100, 76 103 94 200 mM B-ME 1 Values (accurate to i 10% are relative to the control at zero time. Incubations were at 0-2'C. 2 Control incubation buffer was comprised of 5 mM HEPES, 0.1 mM EDTA, pH 7.75. 3 fl-ME, B-mercaptoethanol. 48 this compound (see below). Thiol oxidation may lead to the increase in activity for these samples after 24 h. Kinetic parameters- The rate of urea hydrolysis was monitored as a function of urea concentration from 1 mM to 50 mM urea. Simple Michaelis-Menten kinetics were followed; however, substrate inhibition was observed if the urea con- centration was increased to 100 mM. The data was analyzed by the method of Wilkinson (25) to yield a Kh for urea of 2.8 t 0.6 mM and a.igax of 2800 i 200 umol urea degraded min'l mg'l at 37‘c in 25 mM HEPES, 5.0 mM EDTA buffer, pH 7.75. Some enzyme samples contained B-mercaptoethanol, which was a competitive inhibitor of K. aerogenes urease; however, the concentration of this compound in the assay conditions led to negligible inhibition. pH Studies- K. aerogenes urease retained full activity when assayed under standard conditions following a 30 min incubation at 37°C in buffers ranging from pH 5 to 10.5; however, the enzyme was rapidly inactivated outside of this range. The effect of pH on the kinetic constants for urease was determined in buffers at various pH values. The Kh for urea was nearly constant over the pH range at which the enzyme is stable; i.e., the Kh varied slightly from ca. 2 to 4 mM. In contrast, the “muzexhibited an optimum from pH 7.25 to 8.25. The effect of pH on log igax/Kh is shown in 49 Figure 6. The slope is 1 below the optimum and -1 above the optimum with inflection points at pH values of 6.55 t 0.10 and 8.85 t 0.10 (26). Amino acid analysis- The amino acid composition of K. aerogenes urease is shown in Table 3. Also shown for com- parison are the compositions reported for ureases from B. ammoniagenes (9), jack bean (5), and soy bean (8). The relatedness of K. aerogenes urease to these other proteins was examined by using the method of Cornish-Bowden (27) in which the difference index (DI), the compositional diver— gence (D) and the Marchalonis and Weltman index (SAQ) were calculated. Atomic absorption analysis- Purified K. aerogenes urease was demonstrated to contain nickel by atomic ab- sorption analysis. Three independent preparations of urease were found to contain 2.1 i 0.3 mol Ni/mol 72,000 dalton subunit when compared to a standard curve which was prepared as described in Methods. DISCUSSION The studies reported here extend the work of Friedrich and Magasanik with K. aerogenes urease (13) and of Kamel and Hamed with urease from the related microbe, A. aerogenes 50 i- .- E )6 \ '2' a- X 0 E _ .. :> a 2 -3)- - I- «Ii 4. Figure 6. The effect of pH on log V.“ / K., for K. aerogenes urease. Urease‘Kmu1and K5 values were calculated and the log of the ratio is plotted here in arbitrary units versus the pH of the assay buffer. The buffers include acetate (0), MES (V), HEPES (X), CHES (A), CAPS (o), and phosphate (+). 51 Table 3 Amino Acid Compositions of Ureases Amino Acid K3 aerogenesaB. ammoniagenesbJack bean? Soybeand Cys 1.44 0.38 1.79 0.47 Asx 9.23 11.60 10.60 12.76 Thr 6.90 7.23 6.55 5.28 Ser 6.05 3.64 5.48 5.79 Glx 10.32 10.22 8.22 10.09 Pro 5.50 4.66 5.00 6.09 Gly 12.35 9.68 9.40 10.35 Ala 10.45 10.68 8.80 7.44 Val 7.46 7.80 6.55 6.34 Met 1.68 1.75 2.50 2.04 Ile 5.72 7.13 7.86 6.16 Leu 6.92 7.46 8.21 8.17 Leu 6.92 7.46 8.21 8.17 Tyr 2 1.17 2.5 2.43 Phe 2.57 3.02 2.86 4.23 His 3.00 2.91 2.98 2.09 Lys 3.44 4.87 5.71 5.64 Arg 3.96 4.89 4.52 4.62 Trp 0.99 0.90 0.48 NRe Related in i f D1 0 8.1 9.58 10.62 D 0 0.052 0.057 0.065 SAQ 0 26.8 32.33 41.73 meOU‘m Values represent the mole percent of each amino acid. Taken from (9). Taken from (5). Taken from (8). NR, not reported. Relatedness indices calculated for each urease compared to the K. aerogenes urease using the method of Cornish- Bowden (27). 52 PRL-R3 (14). K. aerogenes urease was purified 1070-fold to homogeneity by standard chromatographic techniques, similar to the procedure used to isolate S. ruminantium urease (11). The final specific activity achieved for the K. aerogenes enzyme (2200 umol ndnfl mg‘l) is among the highest reported for any bacterial urease and approaches that of the jack bean enzyme, about 3500 umol ndnfl mg‘1 or 93 katal/liter [Aum (where a katal is the amount of enzyme which degrades 1 mol of urea/s in a defined pH-stat assay) (1). For com- parison, 24-fold-purified enzyme obtained by Friedrich and Magasanik had an activity of 45.4 umol min'lnrrl (13), and lSO-fold-purified A. aerogenes urease had an activity of 690 umol min‘lrmr4 (14). The Kh for urea of 2.8 mM determined for purified K. aerogenes urease was 4 times that determined by Friedrich and Magasanik of 0.7 mM (13); however, dif- ferent strains were employed. Similarly, the K5 was 2-fold greater with this enzyme compared to the 1.48 mM Km deter- mined for A. aerogenes urease (14). The Vmax calculated for K. aerogenes urease was 2800 umol min‘l mg’l. A novel feature of K. aerogenes urease is the presence of three polypeptide components in the active enzyme. One of these species (hfi.= 72,000) is similar in size to many other microbial ureases, whereas the two smaller components (45.: 11,000 and 9,000) have not been described in any other urease enzyme. Intensive efforts were used to resolve the larger polypeptide from the smaller components; however, the three species remained associated under all conditions 53 which retained urease activity (presence of salt, glycols, thiols, detergents, and mixtures of these substances). These results may indicate that all three polypeptides are subunits of K. aerogenes urease. The three polypeptides are unlikely to be proteolytically derived from a common pre- cursor found in cell extracts, because a variety of protease inhibitors (toluenesulfonyl flouride, bestatin, pepstatin, leupeptin, aprotinin, phosphoramidone, and EDTA) had no ap- parent effect on the SDS-PAGE gel profile for the purified enzyme. The two smaller polypeptides run at the dye front in gels containing less than 10% acrylamide, and these species can be easily overlooked. Indeed, we have found that the S. ruminantium urease, previously thought to possess a singly 70,000-da1ton polypeptide (11), also ex- hibits two small polypeptides on 20% acrylamide gels. Furthermore, in our most highly purified samples of urease from Proteus mirabilis (provided by Julie M. Breitenbach) and Sporosarcina urea (provided by Rick Ye) we have observed a large subunit and two small polypeptides (Figure 5). Thus, one large and two small polypeptides may be generally associated with microbial ureases. Recent studies involving the cloned urease gene from Providencia stuartii support this hypothesis (28). Mobley et al. (28) found that there may be at least two polypeptides peptides (M1. = 73,000 and 25,500) associated with urease in this microbe by using transposon mutagenesis. In contrast to the above microbial enzymes, we have not observed any small polypeptides 54 associated with jack bean urease. Further efforts are clearly required to establish the roles for each of the three K. aerogenes urease components. The quaternary structure and native molecular weight of K. aerogenes urease were not precisely established. From the integrated intensities of the gel scan profiles, an ap- proximate stochiometry for the three components was found to be 1:2:2. This result, combined with the native molecular weight determined by gel electrophoresis and Superose 6 gel filtration chromatography, would indicate a native urease structure (Afi.= 224,000) containing 2 mol of the 72,000- dalton peptide, 4 mol of the 11,000-dalton peptide and 4 mol of the 9,000-dalton component. These results contrast sig- nificantly with the simple homohexameric structure of jack bean urease and the suggested homopolymeric structures (trimer, tetramer, pentamer, and hexamer) reported for other microbial ureases (9, 10, 11, 29, 30). Stability studies of ureases from both jack bean (1) and Arthrobacter oxydans (31) showed irreversible loss of activity below pH 4.5. Similar results were observed with the K. aerogenes enzyme; in addition, a high pH inactivation occurred at pH 10.5 and above. As in the case of the jack bean urease (32), the Kh for urea is nearly constant throughout the entire pH range over which the enzyme is stable. K3 aerogenes urease exhibited a VhaxfiKh dependence on pH which is most easily interpreted by assuming a catalytic requirement for a deprotonated active site group 55 with a pK' = 6.55 and a protonated group with a pr = 8.85. Very early work with jack bean urease indicated similar simple behavior with pk; values of 6.1 and 9.2 (33); how- ever, subsequent studies suggest a more complex behavior (32). In contrast to K. aerogenes urease, the urease from A. aerogenes exhibited very complex pH behavior and was suggested to possess three functional groups with pk; values of 5.6, 6.65, and 8.1 (14). The amino acid composition for K. aerogenes urease was determined and compared to that of jack bean, soy bean, and B. ammoniagenes ureases. K. aerogenes urease was not re- lated to the plant enzymes according to the "weak" test of Cornish-Bowden (27); i.e. the relatedness indices exceeded the recommended limits for a sequence of 840 amino acids. Surprisingly, the two microbial urease compositions also fail to meet the weak test for relatedness; however, the D1, D, and SAQ values indicated that K. aerogenes urease was more closely related to the bacterial than the plant enzymes. K. aerogenes urease has joined the growing list of nickel-containing enzymes which include other ureases, methylcoenzyme M reductase, carbon monoxide dehydrogenase, and certain hydrogenases (35). Thus far, nickel has been shown to be present in ureases from jack bean (6), say bean (8), S. ruminantium (11), B. ammoniagenes (9), B. pasteurii (10), and A. axydans (30). Furthermore, indirect evidence of nickel-dependent growth is consistent with the presence 56 of nickel in ureases from other plants (34) and from S. ureae (31), Aspergillus nidulans (36), Phaeadactylum tri- cornutum (37) and Tetraselmis subcordifarmis (37). The nickel stochiometry is clearly 2 mol nickel/mol subunit in the jack bean and soybean enzymes, whereas either 1 mol (9,10) or 2 mol of nickel (11) per subunit has been found for the bacterial ureases. The K. aerogenes urease results are most consistent with 2 mol nickel/mol of the 72,000- dalton subunit. The environment of nickel and its role in urease hydrolysis have not been determined. 10. ll. 12. 13. 14. References Andrews, R. R., Blakeley R. L., & Zerner, B. (1984) in Advances in Inorganic Biohemistry, (Eichhorn, B. L., and Marzilli, L. G., eds) Vol. 6, pp.245-283, Elsevier Scientific Publishing Co., New York Blakeley, R. L., & Zerner, B. (1984) J3 Mal. Catal. 23. 263-292 Sumner, J. B. (1926) J. Biol. Chem. 69, 436-441 Dixon, N. E., Hinds, J. A., Fihelly, A. K., Gaola, C., Winzor, D. J., Blakeley, R. L., & Zerner, B. (1980) Can. J. Biochem. 58, 1323-1334 Mamiya, G., Takishima, K., Masakuni, M., Kayumi, T., Ogawa, K., & Sikita, T. (1985) Proc. Jpn. Acad. 61, 395-398 Dixon, N. E., Gazzola, C., Blakeley, R. L., & Zerner, B. (1975) J; Am. Chem. Soc. 87. 4131-4133 Alagna, L., Hasnain, S. S., Piggott, B., & Williams, D. J. (1984) Biochem J. 220. 591-595 Polacco, J. C., & Havir, E. A. (1979) J. Biol. Chem. 254, 1707-1715 Nakano, H., Takenishi, S., & Watanabe, Y. (1984) Agric. Biol. Chem. 48, 1495-1502 Christians, S., & Kaltwasser, H. (1986) Arch. .Micrabiol. 145, 51-55 Hausinger, R. P. (1986) J. Biol. Chem. 261, 7866- 7870 Rosenstein, I. J., Hamilton-Miller, J. M., & Brumfitt, w. (1981) Infect. Immun. 32, 32-37 Friedrich, B., & Magasanik, B. (1977) J. Bacterial. 131, 446-452 Kamel, M. Y., & Hamed, R. R. (1975) Acta Biol. Med. Ger. 34, 971-979 57 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29 30. 31. 32. 33. 58 Smith, C. J., Hespell, R. B., & Bryant, M. P. (1980) J. Bacteriol. 141, 593-602 Weatherburn, M. W. (1967) Anal. Chem. 39, 971-974 Lowry O. H., Rosebrough, N. J., Farr, A. L., & Randall, R. J. (1951) J. Biol. Chem. 193. 265-275 Lammeli, U. K. (1970) Nature 227, 680-685 Guilian, G. G., Moss, R. L., & Greaser, M. (1983) Anal. Biochem. 129, 277-287 Eckhardt, A. E., Hayes, C. E., & Goldstein, I. J. (1976) Anal. Biochem. 73. 192-197 Blattler, D. P., Contaxis, C. C., & Reithel, F. J. (1967) Nature 216, 274-275 Fishbein, W. N. (1969) 5V9 International Symposium Chrom. Elect., pp. 238-241, Ann Arbor-Humphrey Science Press, Ann Arbor, MI Zwaan, J. (1967) Anal. Biochem. 21, 155—168 Gurd, F. R. N. (1967) .Methads Ehzymal. 11, 532-541 Wilkinson, G. N. (1961) Biochem. J. 80, 324-332 Siegel, I. H. (1975) Enzyme Kinetics, John Wiley & Sons, New York Cornish-Bowden, A. (1983) .Methads Ehzymol. 91, 60-75 Mobley, H. L. T., Jones, B. D., & Jerse, A. E. (1986) Infect. Immun. 54, 161-169 Carvajal, N. C., Fernandez, M., Rodriguez, F. P., & Donoso, M. (1982) Phytachemistry (Oxf.) 21, 2821- 2823 Creaser, E. H., & Porter, R. L. (1985) Int. J. Biochem. 17. 1339-1341 Schneider, J., & Kaltwasser, H. (1984) Arch. .Micrabiol. 139, 355-360 Dixon, N. E., Riddles, P. W., Gazzola, C., Blakeley, R. L., & Zerner, B. (1980) Can. J. Biochem. 58, 1335- 1344 Laidler, K. J. (1955) Faraday Soc. Trans. 51, 550- 561 34. 35. 36. 37. 59 (1977) Plant Sci. Lett. Polacco, J. C. Hausinger, MacKay, E. M., & Pateman, Microbial. 116, 249-21 Rees, T. A. B., & Bekheet, (Berl.) 156. 385-387 J. A. I. A. R. P. (1987) .Micrabiol. Rev. (1980) (1982) 10, 249-255 51. 22-42 J. Gen. Planta Competitive Inhibitors of Klebsiella aerogenes Urease: Mechanism of Interaction with the Nickel Active Site ABSTRACT We examined several compounds for their mechanisms of inhibition with the nickel-containing active site of homo- geneous Klebsiella aerogenes urease. Thiolate anions compe- titively inhibit urease and directly interact with the met- allocenter, as shown by UV-visible absorbance spectroscopic studies. Cysteamine, which possesses a cationic B-amino group, exhibited a high affinity for urease (K,- = 5 )LM) . whereas thiolates containing anionic carboxyl groups were uniformly poor inhibitors. Phosphate monoanion competi- tively inhibits a protonated form of urease with a pk; of less than 5. Both the thiolate and phosphate inhibition re- sults are consistent with charge repulsion by an anionic group in the urease active site. Acetohydroxamic acid (AHA) was shown to be a slow-binding competitive inhibitor of urease. This compound forms an initial E-AHA complex which then undergoes a slow transformation to yield an E-AHA* complex; the overall dissociation constant of AHA is 2.6 pM. Phenylphosphorodiamidate, also shown to be a slow-binding competitive inhibitor, possesses an overall dissociation constant of 94 pM. The tight binding of 60 nae vVl .aé‘ '“l -4.1 up: AA. vv. Q... 61 phenylphosphorodiamidate was exploited to demonstrate the presence of two active sites per enzyme molecule. Urease contains 4 mol nickel/mol enzyme, hence there are two mol nickel ion/catalytic unit. Each of the two slow-binding inhibitors are proposed to form complexes in which the inhibitor bridges the two active site nickel ions. The inhibition results obtained for K. aerogenes urease were compared with inhibition studies of other ureases and are interpreted in terms of a model for catalysis proposed for the jack bean enzyme (Dixon, N. E., Riddles, P. W., Gazzola, C., Blakeley, R. L., and Zerner, B. (1980) Can. J. Biochem. 58, 1335-1344.) INTRODUCTION Urease (urea amidohydrolase E. C. 3.5.1.5) is a nickel- containing enzyme that hydrolyzes urea to form carbamate and ammonia; carbamate spontaneously degrades to C02 and a sec- ond molecule of ammonia (1-3). The best characterized urease is the enzyme purified from the plant Canavalia ensifarmis (jack bean) (1,2). Jack bean urease is a homo- hexameric enzyme (subunit An = 90,770; 4) which contains 2 mol nickel/mol subunit (5). Zerner and colleagues have pro- posed a model for jack bean urease catalysis in which one nickel coordinates the oxygen atom of urea, polarizing the carbonyl group, and a second nickel coordinates hydroxide ion, the catalytic nucleophile. Three amino acid residues are also proposed to be located at the active site (6): a carboxyl group, a sulfhydryl group, and an unidentified base. Despite extensive investigation the detailed mecha- nism of catalysis has not been established and the structure of the metallocenter is unknown. In contrast to the wealth of studies on the jack bean enzyme, much less in known about catalysis by microbial ureases (3). Clear structural differences from the plant enzyme have been observed; for example, the bacterial urease of Klebsiella aerogenes possesses three subunit types (AA. = 72,000, 11,000, and 9,000) in an 012B474 stoichiometry (7) 62 63 and contains 4 mol nickel/mol enzyme (7). Further charac- terization of microbial urease is important because infec- tious ureolytic microorganisms can contribute to the devel- opment of urinary stones, pyelonephritis, gastric ulcera- tion, catheter encrustation, and other pathogenic conditions (3). In addition, elevated levels of soil microbial urease activity can decrease the efficiency of urea fertilizers (8) and retard seed germination and seedling growth (9). Thus, the study of urease inhibitors may have medical or agronomic significance, as well as providing insight into the urease catalytic mechanism. Here we characterize the inhibition of purified K. aerogenes urease by thiols, phosphate, acetohydroxamate and phenylphosphorodiamidate. We compare these results to pre— vious bacterial urease inhibition studies, and we discuss our results in terms of the catalytic model proposed for jack bean urease. EXPERIMENTAL PROCEDURES Materials- Sources are: acetohydroxamic acid, bovine serum albumin, cysteamine, cysteine methyl ester, B-mercap— toethanol (Sigma); phenylphosphorodiamidate (ICN Biomedicals Inc.); urea (enzyme grade) (U. S. Biochemical Corp.); ethane thiol, 3-mercapto-2-butanol, 2-mercaptopropionate, 3-mercap- tapropionate, methylthioglycolate, 2-propane thiol, 64 3-mercapto-1,2-propanediol (Aldrich); cysteine (Fisher); cresol red (Allied Chemical and Dye Corp.). Enzyme purification- Urease was purified from K. aerogenes (currently K; pneumoniae) strain CG 253 carrying the recombinant plasmid pKAU19 (10) by procedures essen- tially identical to those described previously (7). Since the initial specific activity was much higher in this over- producing organism, the final gel filtration step was not required for maximal specific activity (2500 umol min-1 mg'l). The.Kh'value for urease was reevaluated to be 2.4 mM at pH 7-9 rather than 2.8 mM as reported initially (7). The Kh decreases moderately to 0.72 mM at pH 5-6. The enzyme appeared homogenous by sodium dodecyl sulfate polyacrylamide gel electrophoresis. Enzyme concentrations were based on a molecular mass of 224 kDa (7). Assays- Unless otherwise noted, urease was assayed in the presence of 5 mM EDTA, 50 mM HEPES, pH 7.75 and 50 mM urea at 37'C (standard assay mixture). Reactions were in- itiated by addition of small aliquots ( g 0.1 ml) of enzyme to standard assay mixtures ( 2 2.0 ml) at 37°C. Portions of the assay mixtures were withdrawn at intervals and the ammonia was converted to indophenol (11) which was quanti- tated by measuring the absorbance at 625 nm. Linear regres- sion analysis of the ammonia concentrations versus time gave the initial rate. 65 Alternative assays were used to measure urea hydrolysis when compounds interfered with the indophenol ammonia assay. Cysteine inhibition was examined by using an ammonia micro- electrode (Microelectrodes Inc.). Standard assays with varying urea and cysteine concentrations were run, and the reaction was terminated by addition of 0.1 ml of 2.0 M HNO3 to 0.8 ml aliquots of these assays. A low pH was chosen to terminate the reaction because urease is irreversibly inact- ivated below pH 4.5 and because the ammonia exists as the nonvolatile ammonium ion at low pH values. Ammonia concen- tration was determined immediately after aliquots were made 1.0 M in KOH. Millivolt readings were converted to concen- tration by comparison to an NH4C1 standard. A Spectrophotometric assay was employed to examine the effect of thiourea on urease activity. Urease was added to a 1.0 ml cuvette at 37'C containing 10 mM Tris-HCl pH 7.7, 80 ug/ml cresol red, and varied concentrations of urea and thiourea. Urea hydrolysis was monitored by the increase in the absorbance at 573 nm, which was linear with time between 0.8 and 1.6 absorbance units. Initial rates were calculated by comparison to absorbance changes in a control to which aliquots of KOH were added. Protein was assayed by the method of Lowry (11) with bovine serum albumin as the standard. Spectrascapy- All spectra were obtained by using a Gilford Response spectrophotometer with a 1.0 cm 66 microcuvette (25‘C). Spectral data were transferred to an IBM-PC to calculate difference spectra. Enzyme samples were extensively dialyzed prior to scanning to remove the B-ME present during enzyme storage. RQ.Determinations— For inhibitors which equilibrate rapidly with urease, initial rates were determined in the presence of four different inhibitor concentrations at each of seven urea concentrations (50, 10, 5, 3.3, 2.5, 1.667, and 1.25 mM). Data were plotted according to Lineweaver— Burk (13), and best fit lines were calculated by using the method of Wilkinson (14). Kfi‘was determined from replots of slope or apparent Km values versus inhibitor concentration (15). The response of Rh to changes in pH was investigated by using 100 mM MES (pH 5.00-6.75) or 100 mM HEPES (pH 6.75- 8.50) buffers, with 5 mM EDTA, inhibitor and urea (as indicated above). Inhibitor Binding Rates in the Absence of Substrate- Urease was added to 10 mM EDTA, 100 mM HEPES, pH 7.75 at 37°C containing the indicated concentrations of slow bind- ing, competitive inhibitor. At regular intervals, small aliquots were diluted 40-fold into the standard assay mix- ture, and activities were determined. Linear regression analyses of the initial rates provided correlation coef- ficients r > 0.995, indicating that further inhibitor bind- ing did not occur and that very little enzyme-inhibitor 67 complex dissociated during the assays. Apparent rate con- stants were calculated by linear regression analyses of the data plotted as ln(V,/V0) versus time (V, = initial velocity after t min preincubation with inhibitor and V; = initial velocity of enzyme in the absence of inhibitor). Progress Curves in the Presence of Inhibitar- Progress curves for the rate of urea hydrolysis in the presence of inhibitors were obtained by measuring the ammonia produced as a function of time (Pg. For each experiment, 25-50 time points were taken. A curve-fitting program (see below) was used to fit the data points to: P,=P0+vf:+(V,-- Vf)(1-e*m‘)&;;p (1) where P; = product present at time 0, i? = steady state velocity, VG = initial velocity, and hwp = apparent rate constant (16). Dissociation Rate- The rate of dissociation for the enzyme-inhibitor complex (which was <2% active) was measured by diluting 10,000-fold into an assay mixture at 37'C con- taining the indicated concentrations of urea in 10 mM EDTA, 100 mM HEPES, pH 7.5. The concentration of AHA during the assay was less than 0.1 uM, and the concentration of PPD was less than 0.1 nM. Progress curves for the recovery of 68 activity were monitored for several hours and fit to Equation 1. Dissociation of AHA-urease was also determined for the isolated urease-AHA complex at 37°C, following removal of excess AHA on a Bio-Gel P-6 (0.7 x 14.0 cm) gel filtration column. Dissociation was followed by periodically diluting a small aliquot into the standard assay mixture and com- paring the activity to an appropriate control. The rate of dissociation.(&) was obtained by fitting the fraction ac- tivity recovered as a function of time (Efl to>Equation 2. F,=F0+(I-F0)(1-e"‘) (2) Statistical Methods--Results are provided as the mean : S.D., where applicable. The fitting software (17) is avail- able free of charge to those with a modem. The BASIC pro- gram uses the Marquardt Algorithm (18) to fit by least- squares a set of data points to a curve. The user supplies a function and the partial derivatives with respect to each parameter to be fit. Quantitation of Active Sites- The graphical method of Ackerman-Potter (19) was used to determine the number of active sites per native enzyme molecule. Urease (0.0—10.0 nM) in 100 mM HEPES, 10 mM EDTA, 0.1 mM B-ME, pH 7.75 was incubated with 0, 5, or 10 nM PPD at 0°C. Aliquots were 69 analyzed for activity at 37°C after 2, 24, and 96 h by using the standard assay mixture. RESULTS Thiol Inhibition- Several thiol compounds were shown to be competitive inhibitors of urease, with apparent K; values as provided in Table 1. To determine whether the protonated thiol or the thiolate anion was the actual in- hibitory species, the pH dependence of cysteamine inhibition was characterized (Figure 1). Log Kiresponded linearly to pH changes, with a slope of -1 at pH values below the cyste- amine pk; of 8.1 and with a slope of 0 at pH values above this pkh. This behavior is consistent with the deprotonated thiol acting as the inhibitory species. To compare relative inhibition within the series of thiols in Table 1 the pH dependence of each urease inhibitor should be determined. However, in this initial analysis, the K; values were simply recalculated on the basis of the actual thiolate anion concentrations at pH 7.75 by using published pKh'values. Of the compounds tested, the thiolate with the lowest.K;value (5.0 uM) is cysteamine, which has a positively charged B-amino group at pH 7.75. In contrast, thiol compounds containing an anionic carboxyl group are poor inhibitors at pH 7.75 (e.g. 3-mercaptopropionate and cysteine, with K,- values of >500 uM and 9600 1114, 7O —2 W I I I I I I I I I I I I I I I T T I T I I rI I T I I T] lj I- -1 )- .1 . 1 .. _3_ _. F d A .. .. o- v- . X - . V —4.— _ O) r- 4 O )- -1 _ - -1 - -1 _5_ l h d D d r- -4 1- -1 _6 L 1 1 l 1 n 1 111 1 111 L+ 1+1 LLIJ_L 5.0 6.5 ‘7.0 7515.0 8.5 pH Figure 1. pH dependence of. cysteamine inhibition of urease. The Ki values for competitive inhibition of urease by cysteamine were determined in MES (On!) or HEPES (l-- I) buffers and are plotted as log(K,-) versus pH. 71 Table l Thiol Inhibition of K. aerogenes urease Thiolate Total anion Inhibitor Ref. Thiol K3 K! M. W M cysteamine 0.010 5.0 8.1 20 B—mercaptoethanol 0.55 7.7 9.6 20, 21 3-mercapto—l,2-propane diol 1.7 21 9.66 22 ethanethiol 21 37 10.5 20 cysteine methyl ester 0.12 110 6.53 20 3-mercapto-2-butanol 13.5 190 methylthioglycolate 1.6 850 7.7 2-propane thiol > 100 > 140 10.6 21 cysteine 95 9600 8.7 20 3-mercaptopropionate > 100 > 500 10.05 20, 21, 23 a) Based on total inhibitor concentration regardless of species present at pH 7.75. b) Thiolate anion K,- values at pH 7 .75 are calculated based upon the literature pKa values. 72 respectively). Esterification of the carboxyl group en- hances binding as demonstrated by cysteine methyl ester with a.K;cfi3110 uM. Thiols with uncharged side chains possess intermediate Kivalues ranging from 7.7 uM to 190 nM. The ability of 3-mercapto-2-butanol to inhibit urease indicates that the active site must be accessible to secondary thiols. Spectral interactions of B-ME with urease- In the absence of thiols, K. aerogenes urease displays an absorp- tion spectrum (Figure 2) which includes a peak at 406 nm (e = 2,240 M‘lcm‘l, based on Mr = 224,000). This spectrum is essentially pH-independent (data not shown). Binding of thiolate anions enhances the absorbance in four regions between 800 and 300 nm with the exact wavelength depending upon the thiolate structure. For example, the difference spectra maxima were 322, 374, 432, and 750 nm for B-ME, versus 339, 387, 425, and 750 nm for cysteamine. These spectral interactions are similar to those seen with known thiolate-nickel complexes (e.g. nickel-substituted metal- loenzymes or model compounds; 24). A double reciprocal plot (Figure 3) of the absorbance changes versus B-ME concentration provided the binding con- stant (Ky = 0.38 i 0.12 mM) and the extinction coefficients for the thiolate-nickel complex.(8322,ml= 2230 i 220 M'1 cm-l, 6374 mm = 1060 M-1 cm-l, 8432 m = 530 :t 80 M-1 cm-l). The slight upward curvature at low B-ME concentrations in Figure 3 is a result of the high protein concentration (0.25 73 Absorbonce A Absorbonce l—rTI'UUIUIl‘ VI'TY TIIII'I'I' 300 400 500 600 700 800 Wavelength (nm) Figure 2. UV-visible spectra of urease in the presence of thiol. A, Spectra of urease (56 mg/ml) in 10 mM EDTA, 100 mM HEPES, pH 7.75, at 25°C are shown in the presence of 0.0 mM( ), 0.33 mM (- - . - ), and 5.0 mM (----) B-ME. B, Difference spectra of urease in the presence of 0.2, 0.25, 0.33, 0.5, 1, or 5 mM B-ME were calculated versus the control (0.0 mM B-ME, 56 mg/ml). 74 2513- I (A Absorbonce)" TTTV ii T ..T....,....,...fr....,..... -2£5-—1.5 -{15 GAS 1.5 215 :3f5 415 5L5 ] (ml/1") [B—ME Figure 3. Double-reciprocal plot of llA absorbance versus 1/[B-MB]. The inverse change in absorbance at 25°C was plotted as a function of 1/[B-ME] at A432 nm (0) . A374 m (I). and 1432211,“ (A). 75 mM) required to observe these absorbance changes. A Dixon plot (15), used when the enzyme and inhibitor concentrations are of the same magnitude, yielded the same Kg value (data not shown). The spectrally determined Ky value is identical to the K,- determined kinetically for B-ME inhibition of urea hydrolysis at pH 7.75 (0.55 i 0.08 mM). Phosphate Inhibition- Lineweaver-Burk plots demon- strated that phosphate competitively inhibits urease at pH values from 5.0 to 7.0 (data not shown). Urease is labile at pH values below 5.0 and phosphate inhibition was not exam- ined. Above pH 7.0 phosphate inhibition was not purely competitive, possibly due to the effects of high ionic strength. Within the pH range examined, values of log K} are pH dependent (Figure 4), exhibiting a slope of 1 from pH 5.0 to 6.3 and a slope of 2 from pH 6.3 to 7. Thus, phos- phate inhibition of urease appears to require protonation of two ionizable groups with pk; values of 6.3 and <5. Acetohydroxamate inhibition- Progress curves for urea hydrolysis in the presence of AHA are shown in Figure 5. These curves demonstrate a gradual decrease in urease activ- ity in the presence of AHA. The concentration of both AHA and urea appear to affect the rate of inhibitor binding. These initial results are consistent with slow-binding com- petitive inhibition. Such inhibitors are thought to follow the mechanism shown below (16), where inhibitor (I) competes 76 llJllllllllLl4l '09 (K1) 4 A A 1 l Figure 4. pH dependence of phosphate inhibition. The K,- for competitive inhibition of urease by total phosphate species was determined in MES (O) or HEPES (D) buffers and is shown as log (Ki) versus pH. mM NH4 1.21 :32 . . O l- .” :28 1.0- - . . . - ;2.4 E . 2 0.8. . '.: E 8 I "2.0 II. N H ' . : 2: H I . 1-n 0.51 ‘-1.6 E 1—0-1 53 ‘ A a a) L A A S '3' 0.4- :1'2 “' ;0.6 4" .. . 0.2i {I 'v' .04 . h . 4"€—:--:’---:-1 F 0.0 :11....1.... . .1 1 1 11 1.0.0 Figure 5. Urease progress curves in the presence of AHA. Urease was assayed at 37°C in 10 mM EDTA, 100 mM HEPES, pH 7.75 containing 20 mM urea (solid symbols) or 2 mM urea (open symbols), and 0.0 mM (circles), 1.5 mM (squares). or 3.0 mM (triangles) AHA. The reactions were initiated by addition of urease (with 20 mM urea, [urease] = 140 pM; with 2 mM urea, [urease] = 280 pM), and the product, ammonia, was monitored. The curves were fit as described under “Experimental Procedures." 78 with substrate (S) for enzyme (E) to form an initial enzyme-inhibitor complex (E-I), which is slowly transformed to a more stable complex (E-I*). . It, /. a \ ElizEl’ Scheme I A series of experiments was designed to test this mechanism and to define the rate constants. The effect of AHA concentration on the apparent rate of inhibitor binding in the absence of substrate is shown in Figure 6. At each inhibitor concentration, a first order loss of activity was observed. When the apparent rate con- stants derived from these and other experiments were replot- ted as a function of AHA concentration, they yielded a hy- perbolic curve (Figure 6, inset). Morrison and Walsh (16) have shown that this behavior is expected for an inhibitor which reacts as illustrated in Scheme I. The apparent rate of inhibitor binding (4 follows Equation 3, app ) [I] Ki Ki + l-+ Km where S = substrate, 551:: “2 + LQ/il = Michaelis constant, and K,- = 114/£3. Expanding from this equation, one can show that if 79 O O C) (D (D (D C) d. L "[AHA] (mM) 5 10 U T T l r T 0.0 0.4 'oiar 7,2. 1.6' —4.o. Time (minutes) Figure 6. Rates of AHA binding in the absence of substrate. ln(VflW§) is shown as a function of time for urease (3.5 nM) incubated at 37°C and pH 7.75 in the pre- sence of 0.0 mM (0), 1.0 mM (I), 2.0 mM (A), 3.0 mM (V), 5.0 mM (X), or 10.0 mM (0) AHA. Inset, apparent binding rates are displayed as a function of AHA concentration. 8O [I] £6 << £5 [S] (Eq- 4) [I] + K;(l+ if.) then [I] Bapp ~ £5 [5] (Eq. 5) [I] + K;(l +‘E—') Hence, S l l Ki(l+£K—]- ~ — + "' (Eq. 6) lapp is is [I] and a double reciprocal plot of 1/itc.ipp versus 1/[I] should yield a straight line. The data obtained using AHA exhibit this behavior (Figure 7). The y intercept provides a value fem-£5 = 0.047 r 0.002 s‘1 and the x intercept provides a value for K;==1.4 i 0.2 mM. Figure 7 also shows the result from a similar exper— iment carried out in the presence of B-ME, a competitive inhibitor of urease (see above). B-ME competes with AHA binding; the Ki value (0.26 i 0.09 mM) is in reasonable agreement with the K,- for competitive inhibition of urea hydrolysis by B-ME (K;=:0.55 i 0.08 mM). To demonstrate that AHA competes directly with urea, rate constants for AHA binding at several urea 81 ‘999 . (sec) YT rr fi‘ —1.o —o.5 0.6"015 136 115' '210 __J___. (""WM) [AHA] Figure 7. Double reciprocal plot of llltapp versus 1/[AHA]. Data from Figure 6 (inset) are replotted as 1/11app versus 1/[AHA] (O) . Analogous experiments were carried out in the presence of 1.0 mM B-ME (X). Also shown are apparent rate constants for AHA binding from experiments such as Figure 5 for assays containing 5 mM (0). 10 mM (V), 15 mM (A), and 20 mM (I) urea. 82 concentrations were obtained from progress curves such as those shown in Figure 5. The initial velocity did not appear to be affected by the presence of inhibitor; however, iapp responded hyperbolically to increasing AHA levels at each urea concentration (data not shown), similar to the response of hum seen in the absence of substrate (Figure 6, inset). Double reciprocal plots (Figure 7) show that urea competes with AHA binding, and this chi'1.6 i 0.2 mM is in close agreement with the kinetically determined Kh value of 2.4 i 0.4 mM. The rate of E-I* dissociatirnn 46, was determined by diluting the AHA-urease complex (or an appropriate control) into assay mixtures and measuring urea hydrolysis at 37°C as a function of time (Figure 8). The resulting progress curves (fit to Equation 1 with Vf 2 V0) yielded values for 15 of 9.4 i 1.3 x 10'5 s’1 in 40 mM urea and a.is Ser—O—B(OH); Scheme 7 Thus, a reasonable scheme for urease inhibition by boric acid is formation of the following complex, which is consis- tent with the Zerner model involving nucleophilic attack by hydroxide coordinated to nickel. OH HO - ‘B — O - Ni Scheme 8 OH Such a complex may not be able to form a bridge between the two nickel ions, thus borate does not exhibit slow-binding kinetics. Analogous mechanisms involving nucleophilic at- tack by nickel-coordinated hydroxide on an inhibitor could also be envisioned for formation of the initial AHA and PPD 100 complexes. However, because AHA and PPD are analogs of urea, we favor mechanisms shown in Schemes 4 and 6 for these inhibitors. In summary, the bacterial urease inhibition results are consistent with most aspects of a model previously proposed (6) for the active site of jack bean urease (Scheme 2). The illustrated enzyme-inhibitor complexes are proposed as working models to stimulate further studies. kt) 'I.J 10. 11. 12. 13. Andrews, R. K., References Blakeley, R. L. & Zerner, B. (1984) in Advances in Inorganic Biochemistry (Eichhorn, B. L., Marzilli, L. G., eds.) vol. NY Blakeley, R. & Zerner, B. (1984) 23. 263-292 Mobley, H. L. T. & Hausinger, R. Microbial. Rev., (In Press) Takashima, K., Suga, T. & Mamiya, J. Biochem. 175, 151-165 Dixon, N.E., Gazzola, C., Zerner, B. 4133 Dixon, N. E., Riddles, Blakeley, R. L., & Zerner, Biochem. 58, 1335-1344 Todd, M. J., & Hausinger, Chem., 262, 5963-5967 Mulvaney, R. L. & Bremner, Biochemistry (Paul, Vol. 5. pp. Bremner, Natl. Acad. Sci. USA, P. W., 6, pp. 245—283, Elsevier, J. mol. Catal. P. (1989) G. (1988) Eur. Blakeley, R. L., and (1975) J. Am. Chem. Sac., 97 4131- Gazzola, C., B. (1980) Can. J. R. P. J. M. (1987) J. Biol. (1981) in Soil E. A. & Ladd, J. N., edS) J. M. & Krogmeier, M. J. 153-196, Marcel Dekker Inc., New York (1988) Proc. 85, 4601-4604 Mulrooney, S. B., Pankratz, H. S. (1989) J. Gen..Micrabial. 135, Weatherburn, M. W. (1967) Anal. 971-974 Lowry, O. H., Rosebrough, N. J., Randall, R. J. (1951) J. Biol. 275 Lineweaver, H. & Burk, D. (1934) Sac., 56, 658-666 101 & Hausinger, R. P. 1769-1776 - Chem., 39, Farr, A. L. & Chem., 193, 265- J. Am. Chem. & WC N;- 44 «Ifu Ale ’\ «1C 3 i 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 102 Wilkinson, G. N. (1961) Biochem. J., 80, 324- 332 Dixon, M., and Webb E. C. (1979) in Enzymes, Academic Press Inc., New York Morrison, J. F. & Walsh, C. T. (1988) Adv. Enzymal., 61, 201-301 Schreiner, W., Kramer, M., Krischer, S. & Langsam, Y. (1985) PC TECH J., 3, vol. 5, 170- 190 Marquardt, D. W. (1963) J. Soc. Indust. App. Math., 11, 431-441 Williams, J. W. & Morrison, J. F. (1979) Math. Ehz., 63, 437-467 Danehy, J. P. & Noel, C. J. (1960) J. Amer. Chem. Sac., 82, 2511-2515 DeDekon, R. H., Broekhuysen, J., Bé Chet, J. & Mortier, A. (1956) Bichiama Biophysica Acta 19, 45-52 Sjoberg, B. (1942) Berichte der Deutschen Chemischen Gesellschaft, 75, 13-29 Friedman, M., Cavins, J. F. & Wall, J. S. (1965) J. Amer. Chem. Sac., 87, 3672-3682 Blakeley, R. L., Dixon, N. E. & Zerner B. (1983) Biochim. Biophys. Acta, 744, 219-229 Hase, J. & Kobashi, K. (1967) J. Biochem. (Tbkya) 62, 293-299 Dixon, N. E., Gazzola, C., Watters, J. J., Blakeley, R. L. & Zerner, B. (1975) J. Am. Chem. Sac., 97, 4130-4131 Breitenbach, J. M. & Hausinger, R. P. (1988) Biochem J., 250, 917-920 Dixon, N. E., Blakeley, R. L. & Zerner, B. (1980) Can. J. Biochem., 58, 481-488 Kamel, M. Y. & Hamed, R. R. (1975) Acta Biol. .Med. Germ., 34, 971-979 Nakano, H., Takenishi, S. & Watanabe, T. (1984) Agric. Biol. Chem., 48, 1495-1502 31. 32. 33. 35. 36. 37. 38. 103 Harmon, K. M. & Niemann, C. (1949) J. Biol. Chem., 177, 601-605 Cha, S. (1975) Biochem. Pharm., 24, 2177-2185 Cheng, Y. C. & Prusoff, W. H. (1973), Biochem. 34Pharm. 22, 3099-3108 Blakeley, R. L., Hinds, J. A., Kunze, H. B., Webb, E. C. & Zerner, B. (1969) Biochemistry, 8, 1991-2000 Dixon, N. E., Hinds, J. A., Fihelly, A. K., Gazzola, C., Winzor, D. J., Blakeley, R. L. & Zerner, B. (1980) Can. J. Biochem., 58, 1323- 1334 Neilands, J. B. (1967) Science, 156, 1443-1447 Andrews, R. K., Dexter, A., Blakeley, R. L. & Zerner, B. (1986) J. Am. Chem. Sac., 108, 7124-7125 REACTIVITY OF THE ESSENTIAL THIOL OF KLEBSIELLA AEROGENES UREASE: EFFECT OF DH AND LIGANDS ON THIOL MODIFICATION ABSTRACT The kinetics of Klebsiella aerogenes urease inactiva- tion by disulfide and alkylating agents was examined and found to follow pseudo-first-order kinetics. Reactivity of the essential thiol is affected by the presence of sub- strate and competitive inhibitors, Consistent with a cys- teine located proximal to the active site. In contrast to the results observed with other reagents, the rate of ac- tivity loss in the presence of 5,5'-dithiobis(2-nitroben- zoic acid) (DTNB) saturated at high reagent concentra- tions, indicating that DTNB must first bind to urease be- fore inactivation can occur. The pH dependence for the rate of urease inactivation by both disulfide and alkyla- ting agents was consistent with an interaction between the thiol and a second ionizing group. The resulting macro- scopic pig'values for the two residues are <5 and 12. Spectrophotometric studies at pH 7.75 demonstrated that 2,2'-dithiodipyridine (DTDP) modified 8.5 i 0.2 mol thiol/mol enzyme, or 4.2 mol thiol/mol catalytic unit. With the slow-tight binding competitive inhibitor 104 105 phenylphosphorodiamidate (PPD) bound to urease, 1.1 i 0.1 mol thiol/mol catalytic unit was protected from modification. PPD-bound, DTDP—modified urease could be reactivated by dialysis, consistent with the presence of one thiol per active site. Analogous studies at pH 6.1, using the competitive inhibitor phosphate, confirmed the presence of one protected thiol per catalytic unit. Under denaturing conditions, 25.5 i 0.3 mol thiol/mol enzyme (Afi.=:211,800) were modified by DTDP. I“ 7“ r‘4 INTRODUCTION Urease [urea amidohydrolase (EC 3.5.1.5)] is a nickel-containing enzyme that catalyzes the hydrolysis of urea to ammonia and carbamate; the carbamate spontaneously hydrolyzes to form a second molecule of ammonia and car- bonic acid (2 15). The enzyme from the plant Canavalia ensifarmis (jack bean) is the best characterized urease; however, the role of the plant protein is unclear (29). Dixon et al. (6) proposed a model for the hydrolysis of urea by jack bean urease in which one nickel coordinates the oxygen atom of urea, polarizing the carbonyl group, and a second nickel coordinates hydroxide ion, the cata- lytic nucleophile. Three amino acid residues were sug- gested to participate in catalysis: a carboxyl group, a sulfhydryl group, and an unidentified base. Consistent with the presence of an essential thiol group, both alk- ylating and disulfide reagents were shown to inhibit the enzyme (1, 7, 19). The essential thiol exhibited a pk; of 9.15, and at pH 7.7 reacted more slowly than several other thiols which were not required for activity (19). In contrast to the jack bean enzyme, very little is known about the thiol reactivity of microbial ureases. Characterization of the bacterial enzyme is essential be- cause of its important role in cellular nitrogen 106 107 metabolism and certain pathogenic states (reviewed by Mobley and Hausinger, 15). Ureases from Brevibacterium ammoniagenes (18), Ureaplasma urealyticum (22), and a bo- vine ruminal population (13) have been shown to be inhi— bited by alkylating reagents and p-chloromercuribenzoate; however, the kinetics of activity loss and other proper- ties of the reacting group were not analyzed. We demonstrate that Klebsiella aerogenes urease has an essential thiol and examine the pH dependence of its reactivity with alkylating and disulfide reagents. We also examine the effect of substrate and several competi- tive inhibitors on the rate of inactivation to assess whether this essential thiol is at the active site. Finally, we characterize spectroscopically the reactivity of both essential and non-essential thiol groups in urease . EXPERIMENTAL PROCEDURES Enzyme purificatian- K. aerogenes urease was purified to a specific activity >2500 umol-mindomgd-from strain CG 253 carrying the recombinant plasmid pKAU19 (17) by using procedures previously described (26). For spectroscopic analysis of urease thiols, the final enzyme purification step was carried out with buffers containing DTT rather than B-ME. One unit of urease activity is defined as 1 .1”, v...» o..‘ 9"" at e..- U’ (I) ‘ve Lee ‘1. .M (I) V~ ‘v .‘ :- L 108 umole urea degraded per minute in the standard assay mix- ture, which contained 50 mM urea, 5 mM EDTA, 50 mM HEPES, pH 7.75 at 37°C. Linear regression analysis of the re- leased ammonia, determined by conversion to indophenol (28), versus time yielded the initial rates. Protein was assayed by the method of Lowry (12). Kinetics of urease inactivation- Inactivation re- actions contained 80-100 mM buffer, 8 mM EDTA, plus the indicated reagent concentrations. All thiol modification reagents were dissolved in H53, except DTDP and DTNB, which were dissolved in 100% HPLC grade methanol. Al- though low concentrations of methanol were found to have no apparent effect on urease activity, reaction series using these reagents were kept at a constant concentration of methanol, usually <2%. Inactivation reactions were initiated by addition of a small amount of enzyme (<1% by volume) and aliquots were periodically diluted 50-fold into the standard assay for activity determination. This dilution effectively halted any further inactivation during the assay as shown by the linearity of the initial rate determination. Pseudo-first-order rate constants were obtained from plots of ln(percent activity remaining) versus time. The pH of the inactivation reactions was taken immediately following completion of the assays. Progress curves in the presence of urease inactiva- tors were typically carried out using 5 ml reactions 109 containing 8 mM EDTA, 80 mM HEPES, pH 7.75 at 37°C, and 0.10 ml aliquots were analyzed periodically for ammonia. Experiments in the absence of urea demonstrated a pseudo- first-order loss of initial velocity consistent with Vt : V0 641,th where I; = initial velocity of urease in the absence of modifying reagents, R = concentration of modifying re- agent, and hum = pseudo-first-order rate constant for loss of urease activity. Pseudo-first-order rate constants in the presence of urea were obtained by fitting the product, ammonia, produced as a function of time (Pfi to: vb app (1 .. e—kappt ) (2) PI=P0+k where P0 = product at time t:= 0. Except for DTNB, these rate constants were proportional to the concentration of inactivating reagent at all urea concentrations tested, showing that loss of activity is a second-order process, even in the presence of saturating concentrations of sub- strate (50 mM urea). Kitz and Wilson (9) have shown that the rate of amino acid modification in the presence of ligands can be repre- sented by: (I) (#7 '1 ('3 rr1 Ill (7 'TI rr 110 l '< I E +F? -————€>Inoc1we>enzygm9 2 E L + R ——> Inactive enzyme (Scheme 1) where E = enzyme, R = modifying reagent, L = ligand, (11 = second-order rate constant for the inactivation of free enzyme, ha: second-order rate constant for inactivation of ligand—bound enzyme, and Ki = equilibrium dissociation constant for ligand binding. If the concentration of modifying reagent is much greater than the concentration of enzyme, the effect of the ligand on the apparent pseudo-first-order rate constant, hum, will be given by: it2 (KL+ ,1 [L]) [R] I, <3) app = KL + [L] such that iapp will be dependent on the ratio of £2 to £1: if £2 > (11, the ligand increases kapp; and if (12 < 111, then the ligand decreases hxm' providing complete protection of the susceptible amino acid as (12 approaches zero. Pseudo-first-order rate constants for urease inactivation in the presence of ligands were fit to equation 3 using the Markfit program (see below). 111 Reactivation of DTNB— or DTDP—inactivated urease- Urease was incubated with 2 mM DTNB or with 0.5 mM DTDP for 15 min, then diluted ZOO-fold into 5 mM EDTA, 50 mM HEPES, pH 7.75 at 37°C containing various concentrations of DTT. Aliquots were periodically diluted into the standard assay mixture to measure the activity. Initial velocity data.(l§) were fit to the following expression for the pseudo-first-order recovery of activity: VI: V0 + Vmax (l - e-kappt) (4) where V; = initial velocity of DTDP- or DTNB—modified urease, Vmam = maximum velocity recovered, and kapp = apparent pseudo-first-order rate constant for the reactivation of disulfide-inactivated urease by DTT. Spectroscopic analysis- Phosphate and dithio- threitol, used to stabilize urease during storage, were removed by ultrafiltration using an Amicon Centricon-3O microconcentrator (5 X 1:10 dilution). Spectroscopic data were obtained by using a Gilford Response spectrophoto- meter with l.O-cm.microcuvettes, at 37'C, assuming that 2- thiopyridine had 6343 mm = 7030 M'lcm‘1 (8) . Independent de- terminations of 5u3rm were done using each of the buffer solutions by two methods: aliquots of DTDP were added to cuvettes containing 1 mM B-ME or 1 mM cysteine, or ali- quots of B-ME were added to cuvettes containing 0.5 mM 112 DTDP. Identical results were obtained by both methods, confirming the value of 7030 for 6M3rm of 2-thiopyridine. The total number of urease thiols was determined by re- acting urease with DTDP (0.5 mM) in the presence of 5.5 M guanidine-HCl, 10 mM EDTA, 0.1 M Tris-HCl, pH 7.9. The maximum absorbance at 343 nm was attained within 30 seconds and did not change after an additional 30 minutes. The concentration of protein was determined by the method of Lowry (12), and the molar concentration of urease was calculated based on.Afl.= 211,800 (16). The rates of reaction for the accessible thiols in native enzyme were assessed by scanning three cuvettes for each reaction: a buffer blank, DTDP control, and DTDP + 6—8 uM urease. Spectral data were transferred to an IBM PC to correct for the absorbance increase due to spontan— eous degradation of the thiol reagent (99.9% inhibition. Curve fitting- Curve fitting made use of a BASIC program termed Markfit (23) which employs the Marquardt 113 algorithm (14) to fit by non—linear least squares a set of data points to a rate equation. RESULTS Inactivation of urease by thiol-specific reagents- Homogeneous K. aerogenes urease exhibited a pseudo—first- order loss of activity in the presence of disulfide and alkylating reagents. As illustrated for DTDP, pseudo- first-order rate constants were directly proportional to the concentration of inactivating reagent (Figure 1) allowing calculation of the apparent second—order rate constant. Similar results were obtained for several other reagents, and their second—order rate constants for the loss of ureolytic activity are provided in Table 1. Inactivation of urease by DTNB is also a pseudo— first-order process; however, the apparent rate constant reaches a limiting value at high concentrations of DTNB. Preliminarily, the data were fit to the following model. k k E + R TIE—i E R 4% inocTive enzyme 2 (Scheme 2) The apparent pseudo-first-order rate constant for an inactivator that exhibits a binding step follows 114 c1 -( d -: fl -1 q .1 4 .— .1 .4 .1 ., _ .4 q q q / J. 0.05« Ln < 003-4 Low (9") 0.01- -2.0— , r " O 1 2 ‘ DTDP (mM) 01-1 b I _A O l . 111111111i1111141111111 T T r T r Y T —l T Y I I T f T I Y I 0.0 0.5 1.0 1.5 2.0 Tkne (nfinutes) FIGURE 1. Pseudo-first-order urease activity lose in the presence of DTDP. The fraction of activity remain- ing is shown as a function of time for urease [4.4 nM], reacting with DTDP at 0 (V), 0.25 (O), 0.5 (I), or 1.0 mM (A). All reactions were in 8 mM EDTA, 80 mM HEPES at pH 7.75, and 37'C. INSET: Determination of the second- order rate constant for urease inactivation by DTDP. 115 Table 1 Second—order rate constants for loss of urease activity by thiol-specific reagentsa Inactivator Concentration Rate constant examined 1 Standard error awn) (Mfls-1> Alkylating reagent IAA 10-25 0.002 IAM 1-200 0.026 1 0.001 NEM 0.050—0.200 4.76 i 0.19 Disulfide reagent: DTNBb 0.1-5.0 7.2 DTDP 0.01—4.5 22 i 5 Cystine 0.5-1.0 0.077 t 0.004 Cystamine 0.25-1.0 27 t 1 MMTS 0.025-0.100 710 t 50 a Values determined in 8 mM EDTA, 80 mM HEPES, pH 7.75 buffer at 37'C. Inactivation of urease by DTNB is not a simple second order reaction. The pseudo-first-order rate can be sat- urated at high concentrations of DTNB (Scheme 2); therefore, the apparent second-order rate constant shown 15 M3 k = k2 app 116 i. (3&1 R app _ £1 R + £2 + £3 (5) under steady state conditions (9). Since the inactivation step occurs much slower than the rapid equilibrium between enzyme and modifying reagent, L390% activity retained essentially 100% of the original amount of nickel. Reactivation of urease inactivated by disulfide re- agents- Urease which was inactivated by DTNB slowly recovered activity after dilution into buffer containing DTT, as illustrated in Figure 3. Data for reactivation at three DTT concentrations were fit as described in 117 1300 d I 1 I r I V T I F I I Y T V I I I' I V l 1 be 99 FIGURE 2. Saturation of the apparent rate of urease modification by DTNB: the effects of urea and borate. Pseudo-first—order rate constants for the modification of urease by DTNB were obtained from plots similar to Figure 1. (app-1 (54), obtained at pH 7.75 in 8 mM EDTA, 80 mM HEPES, 37°C, is shown as a function of [DTNB]4-(HWF1) for urease (I), urease in the presence of 2 mM (A) or 5 mM (I) urea. 118 >‘ ;: o 52 T o ai— > 67E “E’EE >. E N c u) 050 [DTT] (mM) O 25 50 75 100 125 Thne (nfinutes) FIGURE 3. Reactivation of DTNB-inactivated urease. Urease (17 nM), previously inactivated by DTNB, was in- cubated in the presence of 8 mM EDTA, 80 mM HEPES pH 7.75 with o (O), 0.1(0), 0.15 (I), or 0.2 (A) mMDT‘I‘ at 37'C. The initial velocity was determined at the indicated times by using the standard assay. INSET: Determination of the second-order rate constant for the reactivation of DTNB- modified urease by DTT. ll9 Experimental Procedures, and the second-order rate constant for urease reactivation (3.1 i 0.2 Ethyl) was calculated from regression of the slope of kum versus DTT concentration. Similar DTT reactivation results were obtained for enzyme inactivated with DTDP (second—order reactivation rate constant 2 2.7 i 0.3 M4snfl. In contrast to the recovery of activity in the presence of DTT, no activity was recovered when DTDP— modified enzyme was treated with excess KCN (500 mM) for 120_minutes at 37'C. Cyanolysis did occur as shown spec— trophotometrically by the release of 4 mole 2-thiopyri- dine/mole enzyme. Addition of DTT (1 mM) to cyanide- treated DTDP—inactivated urease did not release additional 2—thiopyridine. Effect of ligands on the rate of thiol modification- Having shown that thiol-reactive reagents inactivate K. aerogenes urease, experiments were designed to test whether ligands which bind to the active site of urease alter the rate of inactivation. The effect of two competitive inhibitors of urease on the rate of inactivation by IAM is shown in Figure 4. Phosphate was previously shown to be a competitiVe inhibi- tor of urea hydrolysis, with a K}that depended on pH in a complex manner (26). Since phosphate binds with higher affinity to urease at lower pH, the effect of phosphate on urease inactivation was examined at pH 6.1 in MES buffer 120 100 kopp (2'. control) 20a [Competitive inhibitor] (mM) FIGURE 4. Effect of the competitive inhibitors boric acid and phosphate on the rate of urease inactivation by IAN. The second-order rate constants for urease inactiva- tion by IAM were determined in the indicated concentra— tions of phosphate (I) and boric acid (0), and are ex— pressed as percent of the inactivation rate in the absence of inhibitors. Reaction conditions were 8 mM EDTA, 80 mM buffer (MES, pH 6.1 for phosphate inhibition, HEPES pH 7.7 for borate inhibition) at 37'C. 121 where the Kfi,mmqmm£) = 1 mM. Increasing concentrations of phosphate decreased the apparent urease inactivation rate; at saturating phosphate concentrations hum was less than 2% that of the control. Increasing concentrations of boric acid also reduced the apparent rate of inactivation; however, saturating levels of borate decreasedfiéapp by only two-thirds. This difference in behavior between phosphate and borate is not due simply to a pH effect, because borate slowed the rate of urease inactivation by IAM to the same extent at pH 6.0, 7.75, and 9.5 (data not shown). Values for &y%1 (Scheme 1), the ratio of inactivation rates for ligand-bound and free enzyme, and Ki, the ligand dissociation constant, were calculated from the above studies. These values are provided in Table 2, which also lists values obtained in analogous IAM inactivation exper- iments using the inhibitors BBA and 4-Br-BBA and values for the inhibition of urease inactivation using the di- sulfide reagent DTDP. While 4—Br-BBA parallels borate in providing partial protection of the essential thiol, BBA has little effect on urease inactivation by IAM, and ac- tually increased the rate of urease inactivation by DTDP (Table 2). The effect of the ligand urea on the rate of urease inactivation by DTDP was assessed by analysis of progress curves. Urea also provided partial protection of the essential thiol (Table 2), with the apparent second- order rate constant decreasing about 2.5-fold. 122 Table 2 Effect of Urease Ligands on the Rate of Inactivation by Thiol-Specific Reagentsa Inactivator IAM DTDP RQioriKm for urea Ligand hydrolysis flab Ki E2 Ki (mM) 7E1 (mM) "1 (mM) Borate 0.34 0.32 0.30 0.49 0.45 BBA 10 0.8 11 2.6 13 4-Br-BBA 0.37 0.24 0.34 n.d.C ndd. Phosphated 0.8 <0.02 0.50 <0.02 0.43 (pH 6.1) ureae 2.5 n.d. n.d. 0.4 3 a All reactions were done in 8 mM EDTA, 80 mM buffer (pH 6.1, MES; pH 7.75, HEPES) at 37°C. The pseudo-first- order rate of activity loss was measured as a function of inhibitor concentration, then rate constants were fit to equation 3. b £2 and 51 are defined in Scheme 1. n.d. not determined. 6 Initial attempts to fit phosphate data to equation 3 showed (a2 less than 2% (c1, so KL was obtained by fixing ‘12 = 0. e The pseudo-first-order rate for activity loss in the presence of urea was obtained from progress curves in 8 mM EDTA, 80 mM HEPES, pH 7.75, at 37°C and various con- centrations of DTDP. Pseudo-first-order rate constants were obtained by fitting data to equation 2, then these pseudo-first-order rate constants were fit to equation 3. 123 Ligands were also examined for their effects on the rate of activity loss in the presence of DTNB, which was shown above to exhibit saturation of the apparent rate at high reagent concentrations, consistent with Scheme 2. Progress curves for ureolysis in the presence of DTNB yielded rates for activity loss which when plotted as (app4 vs [DTNB]4, were parallel. Saturating concen— trations of substrate decreased the maximum rate, £3, 10— fold, and also decreased the concentration of DTNB re— quired to reach half the maximum rate, kydl, by 10-fold. The effect of pH on t Figure 5 illustrates the app” effect of pH on the rate of K. aerogenes urease inactiva— tion by IAM and DTDP. Above pH 12, the rate does not ap- pear to be affected by changes in pH. Between pH 8.5 and 12, plots of log ham vs pH have a slope of +1. Between pH 8 and 5, the rate of modification is not affected by pH changes; lower values of pH led to rapid denaturation of urease. The pH dependence for DTNB inactivation is comp- licated because of the saturation observed at high DTNB concentrations (as described above). In terms of the re- action illustrated in Scheme 2, the fig/£1 (or K.) of DTNB appeared to be nearly independent of pH, whereas £3 for DTNB modification of urease behaved similarly to the other inactivators (Figure 6). Inactivation by DTNB was exam- ined only below pH 9.5 due to reagent instability. 124 5: Z 0:3 5': 8 I v 4-3 s ”a 32 e 8- : . Q 23 o .. V : DTDP ° ' g 1€OegfleuA :V —l : :1 O—: O, 12 .1 :. IAM I ‘3 :Q..I-AA _2 IIIIIIIIIIIII'IFI] 5.0 7.0 9.0 11.0 13.0 pH FIGURE 5. The pH dependence of IAM and DTDP inac- tivation of urease. The log of the apparent second-order rate constants in (Mfihrl) are shown as a function of pH for DTDP (open symbols) or IAM (closed symbols) for reac- tions carried out in 8 mM EDTA, 100 mM buffer: acetate (circles), MES (squares), HEPES (triangles), CHES (diamonds), CAPS (inverted triangles), or phosphate (circles). 125 I I T I I I T I I I I I I T I I I I I I I I I I I I I T I N O 1 | P’ 01 [L441 :1 I I l 1 Log(Ki or ‘4) L l (L. C) .1 h I T 11 I I l I 9 U1 1 1 II lllllLllllllllllliJllllllJiiJllll 111 I P C) T I I I I I I I T I I T I I l T I I T I 1T7 T I I T T I I ‘10 ‘15 8f) 85 90 95 pH 9’ (n FIGURE 6. Effect of pH and DTNB concentration on the pseudo-first-order rate of urease activity loss. The rate of urease inactivation by DTNB was determined in 8 mM EDTA, 80 mM buffer, at 5 or more concentrations of DTNB at 37’C using MOPS, HEPES, TAPS, and CHES. K,- and k3 (Scheme 2) were determined by non-linear least squares analysis of the pseudo-first-order rate constants versus DTNB concen- tration; log K,- (M) ( O ) and log (c3 (5'1) ( I ) are shown as a function of pH. 126 Spectrophotometric quantitation of urease thiols- In addition to the essential thiol group, other cysteine res- idues may react with disulfides or alkylating agents without loss of urease activity. The total number of urease thiols and the number and reactivity of accessible thiols in native enzyme were probed by Spectrophotometric methods. DTT treated urease (free from low molecular weight thiols) was incubated with DTDP in the presence of 5.5 M guanidine. The number of thiol equivalents was divided by the molar protein concentration to demonstrate that 23.2 i 0.3 mol thiol/mol denatured enzyme could react (assuming the native molecule possesses an 0125474 stoichiometry with Mr = 211,800). Figure 7a shows progress curves for the reaction of DTDP with urease at pH 7.75. The release of 2-thiopyridine corresponded to the reaction of 8.5 i 0.2 mol thiol/mol enzyme. The DTDP—reactive cysteines were converted to mixed disulfides with 2-thiopyridine, rather than to cystine res- idues, as shown by DTT—dependent release of > 8.3 i 0.3 mol 2-thiopyridine/mol of the isolated, inactivated enzyme with concomitant recovery of activity. When urease was equili- brated with the slow, tight-binding competitive inhibitor PPD prior to reaction with disulfide reagents, only 6.3 i 0.4 mol thiol/mol enzyme reacted. Urease modified in the presence of PPD recovered over 70% of its original activity following prolonged dialysis versus 1 mM EDTA, 50 mM phos- phate, pH 6.5, at 4°C, (at 1:20,000 volumes, with several 127 .e O W I!" +5?“MAMA / lllLlllllAll \ / +5opMPPD lllllll I ll] Thiols modified at pH 7.75 N U ¥ 0'! O \J D D an. IIIJJAILLAL AllllllllLJAllllll ._ (D I Q' : "" .2 o d u : .9 r 3‘: '1 U 1 ° 1 E : .2! difference _ .9 1 .c . F‘ : o I I V V I r T U T V r Y T T T I T T V U r T I T r r T Y—r T O 5 1 O 1 5 20 25 30 Time (minutes) FIGURE 7. Spectroscopic progress curves for the re- action of urease with DTDP. Absorbance at 343 nm was con- verted into the number of urease thiols reacting and is shown as a function of time. DTT-treated urease (6-8 HM) was reacted with 0.5 mM DTDP in 8 mM EDTA, 80 mM buffer (HEPES, pH 7.75 or MES pH 6.1) in a 1.0 cm micro-cuvette. Samples include: A) urease at pH 7.75, urease equili- brated with 10 mM AHA or 50 uM PPD, and the absorbance difference between control and PPD inhibited urease, and B) urease at pH 6.1, urease equilibrated with 10 mM phosphate, and the difference scan. 128 buffer changes). In contrast to PPD, in the presence of the slow binding inhibitor AHA, 8.1 i 0.1 mol thiol/mol enzyme were accessible to modification by DTDP, and prolonged di- alysis resulted in recovery of <2% the original activity. Figure 7b illustrates the DTDP reactivity of urease at pH 6.1 in the presence and absence of phosphate. Phosphate protects 2.0 i 0.2 mole thiol/mole enzyme from modification by DTDP. An initial rapid phase was followed by a slow, steady increase (perhaps due to low pH—induced denaturation), so that after 30 min, 8.5 i 0.5 urease thiols had reacted. Figure 8 compares the percent activity of urease as a function of the number of urease thiols modified at pH 6.1 and 7.75 according to the method of Tsou with 1: 1 (27). Tsou plots with c> 1 did not fit the data and are not shown for clarity. The essential thiols reacted more rapidly than non-essential thiols at pH 6.1, possibly as a result of the plateau observed in the pH dependence studies. The slowest phase of urease modification by DTDP at pH 6.1 is not correlated to modification of the essen- tial thiol, as > 95% of the activity is lost by the time 5 thiols have reacted. 129 Fraction Activity Remaining O 03 l # Thiols modified FIGURE 8. Activity as a function of urease modifica- tion. At periodic intervals during the scans in Figure 7, 8.5 pmoles of urease was withdrawn to assay under standard conditions. Fraction activity remaining (aU‘ where 1: 1) is shown as a function of total number of thiols modified at pH 6.1 (I ), and pH 7.75 ( I ), according to the method of Tsou (27). DISCUSSION Demonstration of an essential thiol in K. aerogenes urease- Reaction of urease with either disulfide reagents or alkylating reagents led to at least a 10,000-fold de- crease in activity, consistent with the presence of at least one essential enzyme thiol. The decrease in activ~ ity is probably not a steric effect, since cyanolysis of the bulky 2-thiopyridine group to yield the small un- charged cyanide derivative failed to increase ureolytic activity. Similarly, modification of urease thiols with MMTS, resulting in the small, uncharged methylthio- deriv— ative, led to the loss of >99.9% of activity. Inactiva- tion was not accompanied by a loss of nickel ion, as shown by atomic absorption analysis. Furthermore, DTT-dependent release of either 5-thio-2-nitrobenzoate or 2—thiopyridine led to 100% recovery of the initial activity. Comparison of the second-order rate constants for urease inactivation by the eight thiol-modifying reagents in Table 1 indicates that reagents containing anionic groups react with urease much more slowly than correspond- ing neutral reagents, and the positively charged cystamine reacts over 300-fold faster than the zwitterionic cystine. These observations are consistent with electrostatic ef- fects caused by an anionic group proximal to the essential 130 131 thiol. We previously suggested the presence of an anionic group at the active site of K. aerogenes urease based on comparison of the ngalues for a series of competitive inhibitors (26). These two observations may be related if the essential thiol is at the urease active site. Is the essential thiol at the active site? Loss of urease activity by thiol-specific reagents may be due to modification of an active site thiol, or to modification of a non-active site residue, resulting in a conforma- tional perturbation. In an attempt to discriminate be— tween these two possibilities, the effects of urease ac- tive site ligands on the rate of inactivation was ex- plored. Four competitive inhibitors of urease and the enzyme substrate, urea, were shown to affect the rate of enzyme inactivation, as summarized in Table 2. Whereas urea, boric acid, and 4-Br—BBA afforded only partial protection of the essential thiol, phosphate completely protected the enzyme against inactivation by either IAM or DTDP. In all cases, the effects were achieved at concentrations cor- responding to K}cm'K; of the ligand. These results are consistent with the essential thiol being at the urease active site. Since urea provides only partial protection from inactivation, the essential thiol may not be directly involved in urea binding, but could participate in catal- ysis. These results are consistent with a model for the 132 jack bean urease active site in which a thiol is proposed to serve as a proton donor (6). .pH dependence of urease inactivation by IAM'and DTDP- Roberts et a1. (21) have shown that the intrinsic react- ivity of a thiolate anion with MMTS is >5 X 109 times that of the protonated thiol. Similar results were obtained by Bednar (3) for the reaction of thiols with NEM. Thus, the apparent pseudo—first-order rate constant for either di- sulfide exchange or alkylation can be calculated by Kafi’l "app z [3+] + K“ (7) where %.15 the second-order rate constant for reaction with the thiolate anion, and K; is the acid dissociation constant of the amino acid which is modified. A plot of log(h¥m) versus pH for thiols exhibiting this behavior would possess a slope of 1 below the pK and a slope of 0 above the thiol pKL The pH dependence of urease inact— ivation by IAM and DTDP is inconsistent with this simple mechanism; La the reactivity of the essential thiol is pH independent below pH = 8 for both reagents in Figure 5. One mechanism to explain the data in Figure 5 invokes a greatly enhanced reactivity of protonated reagent to- wards the enzyme thiolate anion. Brocklehurst et al. (4) demonstrated that protonated DTDP (pK = 2.2) reacts with thiolate anions 1500—fold faster than the neutral 133 disulfide. For urease, the equation describing loss of ureolytic activity by this mechanism would be: (app :11 [E-S'] [R] + (2 [E-S'] [R W] (8) with £1 = 105 M‘ls‘l and £2 = 1011 M'lS‘l. This explanation is implausible for DTDP, since kais faster than the dif— fusion rate. A second mechanism consistent with the results in Figure 5 requires the presence of two essential thiols in urease, similar to the mechanism proposed for the reaction of the two active site thiols in thioredoxin reductase (10). In this model, one urease thiol (with a very high pkg = 12) reacts rapidly (Q); a second thiol reacts more slowly (hp and has a low pK} (< 5). The apparent inac- tivation rate constant is given by: K1 £1 [H+] + £2 |H+| I 1 + + -————— K2 [H+] A key point of this model is that inactivation is caused by modification of predominantly one thiol at pH values above 8 and by modification of a second thiol below pH 8. The uniform effect of boric acid on IAM inactivation at pH 6.0, 7.75, and 9.5 appear to preclude this model. If two distinct thiols were reacting at different pH values, one would not expect that a competitive inhibitor would affect 134 the rates equivalently by decreasing the i for modifi— app cation of either thiol by a factor of 3. Our best interpretation of the urease inactivation results is shown in Scheme 3 (modified from Roberts et al., 21), which assumes that the essential cysteine interacts with a second ionizing group (X). C: / \ \<:/.\ inactive enzyme fl': inscttve enzyme (Scheme 3) Two forms of the enzyme can react with IAM or DTDP with second-order rate constants indicated by (:1 and £2. The observed macroscopic equilibrium constants K} and K31 can be related to the individual microscopic equilibrium con— stants of Scheme 3 by equations 10 and 11 (Roberts et al., 1986). K} = K; + K; (10) K3 = K11 + K51 (11) The apparent rate constant for inactivation can then be represented by Itapp = 1 [H+] + KII (12) KI [3+] When the data in Figure 5 were fit to equation 12, values for pKI, pKII, (gig/KI, and £1 were <5, 12, 0.032 M154, and 230 M484, for IAM, and < 5, 12, 11 M454, and 2.0 X 105 M‘ls‘1 for DTDP. Values for £2 alone could not be deter- mined because K3 and K4 (Scheme 3) are unknown. Charact- erization of the individual equilibrium constants would require additional experiments; ag.potentiometric titra- tions of these residues could be carried out analogous to studies involving the active site thiol and histidine of papain (11). Nevertheless, referring to Scheme 3 and the results in Figure 5, the ionization state for X must af- fect the thiol pK such that either pk; approaches 12 when X is deprotonated or pkg approaches 5 when X is protona- ted. An alternative but related model also invokes the presence of a second ionizing group (X); however, here a single intermediate form of the enzyme exists in which the two residues are hydrogen-bonded. The apparent rate con- stant for inactivation can again be represented by equa- tion 9. In summary, the pH dependence of bacterial urease inactivation is most consistent with a single essential thiol interacting with a second ionizable residue 136 (together yielding pk; values of < 5 and 12). By con- trast, Norris and Brocklehurst (19) concluded that the es- sential thiol in jack bean urease has a pk; = 9.15 and does not react with another ionizing group. They examined the essential thiol using DTDP as a Spectrophotometric probe after first blocking the surface thiols with NEM, DTDP or DTNB. Although activity measurements were not re— ported in their pH dependence studies, the spectroscopic reactivity results yielded no plateau in the low pH end of the log rate versus pH plot, down to pH 7.0. However, the jack bean urease study was carried out using phosphate buffers below pH 8. Figure 4 illustrates that phosphate decreases the rate of inactivation of the bacterial enzyme. Furthermore, we have shown (26) that phosphate inhibition is pH dependent, with phosphate becoming a more potent inhibitor as the pH decreases. Thus, a low pH plateau in the jack bean study may have been masked by phosphate-dependent thiol protection. Urease inactivation by DTNB- In contrast to the simple, second-order reaction observed between urease and six of the reagents examined, the pseudo-first—order rates of inactivation were not proportional to DTNB con- centration. Rather, the data were fit to Scheme 2, invol- ving a preliminary DTNB binding step prior to disulfide exchange leading to inactivation. This type of saturation behavior is not uncommon (9); indeed similar kinetics 137 could occur for the other reagents, but the saturating concentrations are higher than the range examined. The pH dependence of urease inactivation by DTNB is consistent with a modified Scheme 3 which includes reversible forma- tion of a complex between enzyme and modifying reagent. Binding is unaffected by the ionization status of the es- sential thiol, whereas the DTNB inactivation step behaves similarly to amp in urease reactions with IAM or DTDP. The effect of urea on the rate of urease modification by DTNB can be accounted for, in part, by a modification of Scheme 1 to yield Ki k E + R \ \ E R -—-‘—> inactive enzyme +L +L KL L Ki k E L + R \ \ EL R ——> inactive enzyme (Scheme 4) As shown by Cardemil (1987), the rate of inactivation in this case is 1R1[KL11 Ki +12 1L1 Ki] kapp : u u (13) Ki KL (Ki + [R]) + Ki [L] (Ki + [R]) The observed parallel lines in the l/Qmp'versus 1/[DTNB] plot (Figure 6) require that 138 it1 Ki it2 K; so that-the binding affinity increases (Ki decreases) as the rateefiz decreases with increasing concentration of ligand. We do not have an explanation for this fortuitous relationship. Spectroscopic quantitation of urease thiols- Using subunit molecular weights and cysteine contents derived from recently published sequence data (16) and the (1213474 subunit stoichiometry derived form integration of bands in Coomassie Blue stained SDS gels (25), the holoenzyme is predicted to possess 20 thiols. Modification of dena- tured enzyme by DTDP showed that urease possessed 25.5 i 0.3 mol thiol/mol enzyme based on Lowry protein deter- mination. These results are consistent with an absence of disulfides in the native protein. Spectroscopic an- alysis of jack bean urease using DTNB combined with the amino acid composition indicated that 15 thiols and one disulfide were present per subunit (Afi.= 96,600; 20), whereas protein sequencing later demonstrated the pre- sence of only 15 cysteines per subunit (Nfi.= 90,770; 24). We have previously shown that K. aerogenes urease possesses two active sites per native molecule (26). hence, we can conclude from the spectroscopic analysis 139 that 4 thiols per catalytic unit react with DTDP at pH 7.75. Modification of at least one of the these thiols is associated with urease inactivation. In the presence of PPD, a tight-binding competitive inhibitor of urease, only 3 DTDP reacted per catalytic unit, and the enzyme was active after PPD removal. These results are consis— tent with the presence of a single essential thiol per active site. AHA, a slow binding inhibitor, did not pro— vide comparable protection of the essential thiol. Studies at pH 6.1 using the competitive inhibitor phos- phate also demonstrated that inactivation was associated with the modification of a single thiol per catalytic unit. 10. 11. 12. 13. 14. 15. References Andrews, A. T., & Reithel, F. J. (1970) Arch. Biochem. & Biophys. 141, 538-546 Andrews, R. K., Blakeley, R. L., & Zerner, B. (1988) in The Bioinorganic Chemistry of Nickel (Lancaster, J. R., Jr, ed) pp. 141-165, VCH Publishers, Inc., New York Bednar, R. A (1990) Biochemistry 29. 3684-3690 Brocklehurst, K., Stuchbury, T., and Malthouse, J. P. G. (1979) Biochem. J. 183. 233-238 Cardemil, E. (1987) in Chemical.Mbdification of Enzymes (Jaime Eyzaguirre, ed) PP 23—34, Ellis Horwood Limited, West Sussex, England Dixon, N. E., Riddles, P. W., Gazzola, C., Blakeley, R. L. & Zerner, B. (1980) Can. J. Biochem. 58. 1335—1344 Gorin, G. & Chin, C—C. (1965) Biochim. Biophys. Acta 99. 418-426 Grassetti, D. R., & Murray, J. F. (1967) Arch. Biochem. Biophys.119. 41-49 Kitz, R. and Wilson, I. B. (1962) J. Biol. Chem. 237. 3245-3249 Kallis, G-B & Holmgren, A. (1980) J. Biol. Chem., 255, 10261-10265 Lewis, S. D., Johnson, F. A., & Shafer, J. A. (1976) Biochemistry 15. 5009-5017 Lowry, O. H., Rosebrough, N. J., Farr, A. L. & Randall, R. J. (1951) J. Biol. Chem.193. 265-275 Mahadevan, S., Sauer, F. D., & Erfle, J. D. (1977) Biochem J. 163. 495—501 Marquardt, D. W. (1963) J. Soc. Indust. App. Math. 11. 431-441 Mobley, H. L. T. & Hausinger, R. P. (1989) .Microbiol. Rev. 53. 85-108 140 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. Mulrooney, S. M., 141 (1989) J. Gen. Microbiol. 135. 1769—1766 Mulrooney, S. M., 172. 5837-5843 Nakano, H., Takenishi, S Agric. Biol Chem. 48, 1495—1502 Norris, R. & Brocklehurst, K. (1976) 159. 245-257 L., Biochem J., and Olson, Riddles, P. W., Andrews, R. K., Blakeley, R. Zerner, B. (1983) Biochimica et biophysica Acta, 743. 115-120 Roberts, D. D., Lewis, S. D., Ballou, D. P., T. & Shafer, J. A. (1986) Biochemistry 25. 5601 Saada, A.-B. & Kahane, I. (1988) Zentralbl. Bakteriol. Parasitenkd. Infektionskr. Hyg. Abt. Orig. Reihe A 269, Schreiner, W., (1985) PC Tech. J. 3. Takishima, K., Biochem. 175. Tsou, C. L. (1962) Kramer, M. Suga, T., 151-165 Todd, M. J. & Hausinger, Chem. 262. 5963-5967 Todd, M. J., & Hausinger, R. P. (1989) Chem. 264. 15836-15842 Weatherburn, M. Winkler, R. G., Randall, D. D. 100 Blevins, 160-167 & Mamiya, G. (1988) 1 5595- Scientia Sinica 11, 1535-1558 R. P. (1987) J. Biol. J. Biol. W. (1967) Anal. Chem. 39. 971—974 D. G., Polacco, J. C., & Trends Biochem. Sci. 13. 97- (1988) Pankratz, S. & Hausinger, R. P. & Hausinger, R. P. (1990) J. Bact. . & Watanabe, Y. (1984) S. , Krischer, S. & Langsam, Y. vol. 5, 170-190 Eur. J. IDENTIFICATION OF THE ESSENTIAL CYSTEINE RESIDUE IN RIJHESIEHHM!.AEHHWSENEIIIHUEASE .ABSTTUKEP During reaction with l4C—iodoacetamide at pH 6.3, radioactivity was incorporated primarily into a single K. aerogenes urease peptide concomitant with activity loss. This peptide was protected from modification at pH 6.3 by inclusion of phosphate, a competitive inhibitor of urease, which also protected the enzyme from inactivation. At pH 8.5, several peptides were alkylated; however, modification of one peptide, identical to that modified at pH 6.3, paralleled activity loss. The N-terminal amino acid sequence and composition of the peptide containing the essential thiol was determined. Previous enzyme inactiva- tion studies of K. aerogenes urease could not distinguish whether one or two essential thiols were present per active site (20); we conclude that there is a single essential thiol present and identify this residue as Cyslm in the large subunit of the heteropolymeric enzyme. 142 INTRODUCTION Urease (urea amidohydrolase, EC 3.5.1.5) is a nickel- containing enzyme that catalyzes the hydrolysis of urea to form ammonia and carbamate; carbamate spontaneously hydro- lyzes to H¢CO3 and a second molecule of ammonia. The best characterized urease is the homohexameric protein from Canavalia ensiformis (jack bean) which contains 2 nickel per subunit (NQ.= 90,770) (reviewed by Andrews et al., 2). The function of urease in this and other plants is unknown (21). In contrast, bacterial ureases play important roles in ni- trogen metabolism and have been implicated as virulence factors in various human and animal diseases (reviewed by Mobley and Hausinger, 9). The most extensively studied bacterial urease is that from the Gram-negative enteric bacterium, Klebsiella aerogenes (currently K. pneumoniae). This enzyme possesses three different subunits [Nfi.= 60,304 (a), 11,695 ([3), and 11,086 (7)], in an apparent a2fl4‘y4 stoichiometry, and has two bi-nickel active sites (18, 19). The DNA-sequence of the K. aerogenes urease operon (11) re- vealed that each bacterial subunit possesses significant ho- mology to portions of the plant urease sequence, consistent with gene fusion in the later case. As described below, both of these enzymes possess an essential cysteine residue; 143 144 however the bacterial and plant enzyme thiols differ in their chemical properties (13; 20). Labelling studies of jack bean urease indicated that 4 cysteines/subunit react rapidly with a disulfide reagent, DTDP, with no loss of activity, whereas modification of a fifth cysteine was correlated to activity loss (13). The essential thiol exhibited a pH dependence of reactivity similar to that of a free thiol, with a pKh = 9.1. Taki— shima et al. (16) followed up on this work by modifying the non—essential enzyme thiols with N-ethylmaleimide, then derivatizing the essential thiol with the chromophore N-(4— dimethylaminodinitrophenyl) maleimide. They demonstrated the specific labelling of a single cysteine residue, Cyssmp In contrast, Sakaguchi et al. (15) partially modified jack bean urease with diazonium-l H-tetrazole, then added 1%}- 1abelled reagent to derivatize the more slowly reacting thiols. They showed that label was incorporated into 2 cyanogen bromide fragments; one fragment was further studied and contained Cysmn. More recently, jack bean urease was modified using N-iodoacetyl-N'(5—sulfo-l—naphthyl)ethylene diamine and N-(7-dimethylamino-4—methyl-3—coumarinyl)male- imide (17): three thiols were modified with no loss of ac- tivity, whereas alkylation of 2 additional thiols (Cyszv7 and Cysan) was concomitant with activity loss. Complete modification of Cysun with the latter reagent resulted in an activity decrease of only 50%; thus the authors concluded that CYSgn (but not Cysfln) is essential for activity. 145 No active site labelling studies have been reported for any bacterial urease; however kinetic analysis of K. aerogenes urease revealed a pseudo-first-order loss of ac- tivity in the presence of disulfide reagents or alkylating reagents (20). The addition of substrate or inhibitors altered the rates of inactivation, consistent with local- ization of the essential thiol proximal to the active site. Competitive inhibitors (PPD at pH 7.5, and phosphate at pH 6.1) were shown to protect one thiol per active site from modification by DTDP. The apparent second-order rate con- stant for activity loss as a function of pH (using either IAM or DTDP) differed notably from the behavior of the plant enzyme in being constant over the range of pH 5 to pH 8. The results are consistent either with the essential thiol interacting with a nearby ionizable amino acid or with the presence of two essential cysteines per active unit (20). In the latter case, inactivation at low pH would involve modification of a cysteine which is distinct from that modified at high pH; both cysteine residues would be required for activity. This manuscript describes our efforts to identify the essential thiol(s) of K. aerogenes urease and to discrimi- nate between the one- and two-cysteine models. EXPERIMENTAL PROCEDURES Enzyme Purification— K. aerogenes urease was purified to a specific activity >2500 umol-min*bn@“1 from strain CG253 carrying the recombinant plasmid pKAU19 (12) by using procedures previously described (19). The third and final step of the purification, Mono-Q (Pharmacia) anion exchange chromatography, made use of DTT-containing buffers (1 mM) to ensure the absence of mixed-disulfides between urease and B- ME. Enzyme was stored at O’C in DTT-containing buffers un- til used. Protein was assayed by the method of Lowry et al. (8) using bovine serum albumin as the standard and confirmed using the extinction coefficient 8:30“ = 8.5 (18) . The mo- lecular weight of urease was calculated to be 211.8 ug/nmol (11). Uniform labelling of cysteine residues in urease- Purified urease was desalted by ultrafiltration using a Cen- tricon-30 microconcentrator (5 X 1:10 dilution) to remove thiols used to stabilize the enzyme during storage, then dried using a speed-vac concentrator (Savant). 15 mg urease was denatured in 6 M Gn-HCl, 10 mM EDTA, 0.1 M Tris-HCl, pH 8.0, reduced by incubation with DTT (500 nmole) for 60 min at 50°C, and carboxamidomethylated with 14C-IAM (2.5 mole @ 0.185 GBq/mmol, Amersham) for 30 min at 25°C. The modifica- tion reaction was quenched by adding excess B-ME. 146 147 Specific labelling of the cysteine residues in native urease- K. aerogenes urease (20-50 nM) was desalted into 1 mM EDTA, 80 mM buffer (MES, pH 6.3, HEPES, pH 7.75, or TAPS, pH 8.5) and allowed to react with 5 mM 1NZ-IAM at 37°C until < 5% of the activity remained, as determined by assaying aliquots using a standard assay procedure (19). Reactions at pH 6.3 were done both with and without 20 mM phosphate, a pH-dependent competitive inhibitor of urease (19). to pro- tect the essential cysteine from modification. Some reac— tions at pH 7.75 included the competitive inhibitors AHA (25 mM) or PPD (25 nM). During the reactions, aliquots were quenched by addition of excess B-ME, then desalted on a Superose-12 gel filtration column (1 X 30 cm, Pharmacia), equilibrated with 1 mM EDTA, 20 mM phosphate, pH 6.5, con- taining 1 mM B-ME. Radioactively labelled urease samples were denatured in 6 M Gn-HCl, reduced with DTT, and the re- maining cysteines were modified with unlabeled IAM using procedures analogous to those described above. Standardized protocol for peptide mapping- The a sub- unit of urease was separated from the B and Y subunits by Superose-12 gel filtration chromatography (1.0 X 30 cm) in 3 M Gn-HCl, 1 mM EDTA, 0.1 M Tris-HCl, pH 8.0. Fractions con- taining the a subunit were pooled, concentrated, and ad- justed to 1 M Gn-HCl. CaClZ was added to 1.0 mM, then 148 samples were digested with 4% TPCK-treated trypsin (TYpe XIII, Sigma) for 24-40 hours at 37'C (2% added initially, another 2% added after 8—12 hours digest). Tryptic peptides were separated by using reverse phase chromatography (Pro- RPC, 0.5 X 10.0 cm, Pharmacia), with a multistep gradient of 0—50% IPN containing 0.05% TFA. Two peptides which did not bind to the Pro—RPC could be separated on a Pep-RPC column (Pharmacia) with a 0-100% acetonitrile gradient, containing 0.1% TFA. Characterization of the specifically labelled peptide- Sequence analysis of the specifically-labelled tryptic pep- tide was done by using an Applied Biosystems Model 477 pro- tein sequencer followed by HPLC analysis of phenylisothio— cyanate derivatives. Amino acid composition was determined by HPLC analysis after converting the hydrolyzed sample to the phenylisothiocyanate derivatives. RESULTS AND DISCUSSION Uniform labelling of urease cysteines and development of a peptide mapping protocol- Denatured K. aerogenes urease was uniformly labelled with 14C-IAM at pH 8.0. The a subunit (M1. = 60,304) was separated from the B and 7 sub- units (Afl.= 11,695 and 11,086, respectively) by gel-filtra— tion chromatography (Figure 1), which also removed excess 149 I Tl ' ' I I T I F I t r r T 5000: 30001 cpm I 10005 I I j __ .---____I -_ L==__JL..=_J L I a 5+7 Excess Reagent 0.4— 280nm 0.2a 0.0 ,....r,...?..A/{\.:~., 5 to 15 20 25 Vdunm OnD Figure l. Subunit distribution of urease thiols. Urease was denatured, reduced, and uniformly modified with l4C-IAM. Alkylated protein was applied to a Superose-12 column (1.0 X 30 cm, Pharmacia) equilibrated with 3.0 M Gn~HCl, 10 mM EDTA, 0.1 M Tris-HCl, pH 8.0. Cpm was determined in 5% of each 1.0 ml fraction, and absorbance was monitored at 280 nm. Brackets indicate fractions included in calculations of the relative amount of label in each subunit. 150 reagent. The ratio of 14C in the a subunit to that in the sum.of the B and 7 subunits was 4.8:1. The B and 7 subunits could be separated by reverse phase chromatography (data not shown), and all of the radioactivity was found in the B sub- unit; thus the ratio of thiols in the a, B, and 7 subunits was 4.8:lzo. The genes encoding these proteins (11) indicate an 8:1:0 ratio of thiols in the a, B, and 7 subunits. Hence, there are twice as many B as a subunits in the native enzyme, consistent with the proposed a2B4y4 subunit structure (18). Purified carboxamidomethylated-urease a subunit was highly insoluble unless maintained in the presence of a de- naturant such as l M Gn-HCl. A standard protocol for pep- tide mapping was developed by digesting the subunit dis— solved in this chaotropic agent with TPCK-treated trypsin for various lengths of time and analyzing the peptides by Pro-RPC chromatography (Figure 2). Optimum resolution of cysteine-containing peptides (determined by monitoring ab- sorbance at 214 nm and by analysis of fractions for cys- teine-containing peptides using liquid scintillation coun- ting) was only achieved after 30—40 hours of digestion, at 37'C. Six major cysteine-containing peaks were identified as flow-thru (FT), and peptides A, B, C, D, and E (Figure 2). The flow-thru fraction could be further resolved into two additional peptides, F and G, on the more hydrophobic column, Pep—RPC (data not shown). This protocol yielded consistent peptide maps in which the 7 peaks accounted for 151 800- - W 500 A 1:E 50 C 5: F D i _ Cpm) 400 . 3 40 200 V E (30 % IPN 0-44 o...---‘~2o o.3« . Absorbonce ’,' 214 nm 0.2- .10 0.1. W“ MN 0.0 .. W, o 50 60 Volume (ml) Figure 2. Pro-RFC analysis of uniformly-labelled trypsin-digested urease arsubunit. Purified urease a- subunit was digested with TPCK-treated trypsin (as described in the text) and applied to a Pro-RPC column. The column was eluted using the indicated gradient of IPN/0.05 % TFA (dotted line) while monitoring the absorbance at 214 nm (solid line). Fractions (1 ml) were collected and radio- activity was determined in 0.1 ml aliquots by scintillation counting (dashed line). 152 at least 75 % of the applied label in uniformly-labelled samples, and clearly resolved the specifically labelled pep- tides in the experiments described below. Specific labelling of the essential cysteine at pH 6.3— The reactivity of the essential thiol of K. aerogenes urease exhibited a pk; of 12 and decreased below this value until about pH 8; below pH 8 the essential thiol showed a constant rate of modification (20). In contrast, the reac- tivity of most thiols towards modifying reagents decreases with decreasing pH below the thiol pkg. We attempted to exploit this low pH reactivity as a method to specifically label the essential thiol. In addition, since the competi- tive inhibitor phosphate was shown to completely protect urease from inactivation by both disulfide-forming reagents and alkylating agents (20), we sought to identify the essen— tial cysteine modified at pH 6.3 by comparing the peptides labelled by lflt-IAM in the presence and absence of phos- phate. After 25 hours incubation with L4C-IAM in the ab- sence of phosphate, a total of 7.7 thiols (6 in the a sub- units, l.7 in the B-subunits) were modified with complete loss of activity; in the presence of phosphate, 4 thiols in the a subunits were protected from modification (Figure 3). Phosphate did not protect any cysteine in the B subunit, demonstrating that the essential cysteine is in the a sub- unit. The finding that only 2 B-subunit thiols are modified in inactivated enzyme may indicate that the 4 B-subunits are 1 o I U V I I I V l T I I _____ o- _ J 0.3- . Fraction . O O 006‘ C1 Actwnty Remaining 0.4- 4 (L2- ITtIITTTWIT‘rTIIIYTIYUTT' 1 65% 5.0- Number of . Thiols ‘ Rea cted 3'0: 2.0- VT—TITrTTII" YT' I'l'i" o 5 10 '1'5‘ 20 25 Time (hours) Figure 3. Effect of phosphate on the modification of urease thiols by IAM. Urease was inactivated with 5 mM LR}- IAM in the presence (open symbols) and absence (closed symbols) of 20 mM phosphate, in 1 mM EDTA, 80 mM MES, pH 6.3. A) The remaining activity is shown as a function of time. B) The number of a-subunit (squares) and B-subunit (triangles) thiols reacted was assessed as described in the text. 154 present in 2 distinct conformations: the thiols in two of these subunits are accessible to solvent, whereas the other 2 B-subunit thiols are sequestered. Purified a—subunit was digested with trypsin, and sub- jected to the Pro-RPC peptide mapping protocol. In the ab- sence of phosphate, the label was incorporated most specifi- cally into peptide D, and the extent of peptide modification was reasonably well correlated to activity loss. Further- more, phosphate provided significant protection of this same cysteine-containing peptide (Figure 4). Other cysteine-con— taining peptides were protected to a lesser extent, probably because phosphate stabilizes the enzyme from low-pH induced denaturation. Specific labelling of the essential cysteine at high pH- At higher pH values, competitive inhibitors of urease did not protect the essential thiol from alkylation by IAM (Figure 5). This result was surprising because we had pre- viously shown that the slow-tight binding inhibitor PPD (K} = 95 pM), but not AHA (K}=:2.6 pM), protected 1 mole of thiol per mole of active site from modification by DTDP (20). PPD may be able to sterically exclude DTDP from the active site, whereas the smaller, uncharged alkylating re— agent, IAM, is still able to react with the essential thiol. Because the competitive inhibitor protection approach was untenable for identification of the essential cysteine 155 ‘LO 7‘ (LB- 3 ‘ 10hr “hr 2°” 3‘3"? o 6- ‘00 ' OI E Q 7 U) 3§15 Ouk- ll 5 I 4—1 d y 5535" , 355' . SSN ’5 . I :I'IN :II 'i'1°§ 7 {4 ’//I .;.;.§ .;.;.I o 0 25:3 1:1 %. E125! 52:38) 32:3 k? 1:39 Izt Figure 4. Peptide distribution of urease thiols modified at low pH. The a-subunit of urease, modified at pH 6.3, was purified and digested with trypsin. Pro—RPC chromatography was used to determine the amount of labelled thiol in each peptide in the presence (+) and absence (-) of phosphate. 156 Y T IIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIIII 8 l I I I d )- 1004? . \ ‘\ l e d “ . ao— ‘ *5 . 'o 0.) .4 “ E § " 'o eo~ \ ,//‘ ~ g .. ‘v % ActIVIt « \ c4 y ~ 1’ remamm ‘ a a. 9 9 40— *x -c . / cg _ r- q “\..~‘ _ 2 % 20"I “~‘.‘.. )- ........... .---- u- - ________ .u 0 IIIIIIIII I I rrrrrrrrr T T TTTTT I ''''''' T T r 1 WT T T I F 0 O 5 10 15 20 25 Tune (hours) Figure 5. Effect of competitive inhibitors on urease thiol modification at pH 7.75. Urease was inactivated at pH 7.75 in 10 mM EDTA, 100 mM HEPES, pH 7.75 using 14C-IAM, in the absence of inhibitors (0), and in the presence of 25 mM AHA (I) or 25 MM PPD (A) . Aliquots were quenched with an excess of B-ME, desalted using Superose-12 gel-filtration chromatography, and the incorporated radioactivity was measured to determine the number of thiols modified. The time-dependent decrease in activity is that for urease inactivated in the absence of inhibitors (4»). Activity of the inhibited samples was not determined because the rate of inhibitor dissociation to yield active enzyme is slow (T05 ~ 2.5 hr, Todd and Hausinger, 1989). 157 residue at higher pH values, the reactivity of individual cysteine-containing peptides was examined in uninhibited samples as a function of remaining activity. The pattern of subunit labelling at pH 8.5 is shown in Figure 6a. 2-3 thiols were modified with essentially no loss of activity; activity loss occurred upon modification of an additional 5- 6 thiols per halo-enzyme, with 1-2 thiols modified in the B subunits, and 4 thiols modified in the a subunits. Tryptic peptide analysis of a-subunit purified from urease modified at pH 8.5 revealed that activity loss paralleled the labelling of only peptide D (Figure 6b). As in the pH 6.3 study, only half of the B-subunits were alkylated at pH 8.5, perhaps indicating that the halo- enzyme has two distinct populations of B-subunits. (We assume that the same two thiols are reactive at each pH). The low pH/inhibitor protection analysis demonstrated that the B-subunit thiol is not essential for catalysis; however, modification of this thiol (at pH 8.5) closely followed ac- tivity loss in the absence of inhibitors. The observed similarity in reaction rates between the B-subunit thiol and the essential thiol in K. aerogenes urease may relate to the anomalous jack bean urease active site labeling studies of Sakaguchi et al., (15). The bacterial B-subunit is homolo- gous to residues 132-237 in the jack bean urease sequence (11, 17) and the only cysteine in this region of the plant urease is Cysmfl; Sakaguchi et al. (15) may have 158 4 4 O.9~ . . 0.7- - Fraction ' Actwuiy. 0.5« J Remaining ‘ a3~ . i 1 «i OJ- . to a .. ‘\ 5 i «'5 ‘ I.A o.s- ‘ B e C 4 ' D . a Fraction 0-6- . E + G Acfivhy i \ Remaining 0.4- K \ QO . , . . . , . an 02. 0A- at as Thiols modified Figure 6. Peptide distribution of urease thiols modified at high pH. Urease was inactivated with 5 mM 1%}— IAM at pH 8.5. A) The activity remaining as a function of the total number of thiols modified (0) was determined. In addition, the distribution of label in the a (I) and B (A) subunits was measured. B) Activity is shown as a function of the amount of label incorporated into the a-subunit tryptic peptides. 159 misidentified Cysun as the essential cysteine if this residue reacts similarly to that in the bacterial B-subunit. In contrast to jack bean studies reporting the rapid modification of several thiols followed by slower reaction with the essential thiol (13, 1516, 17), we saw no cysteine- containing peptide in the a-subunit labelled faster than peptide D. Rather, our peptide mapping protocol demonstra- ted that a small amount of each cysteine-containing peptide was labelled with no loss of activity, perhaps due to the presence of a small (<10%) fraction of inactive protein. Identification of the cysteine in peptide D- N- terminal sequence analysis of peptide D, purified from uniformly labelled urease, revealed the sequence TLNTI. Comparison of this sequence to that predicted from DNA sequence analysis (11) allows us to conclude that peptide D begins at residue 305 of the a—subunit. This peptide is produced by cleavage after Tyrum, indicating that after 30 hours digestion, significant Whtryptic cleavage had oc- curred. WLtrypsin is an autolytic product of trypsin which possesses chymotryptic-like activity, although at a much slower rate (1). Peptide D was subjected to amino-acid analysis (Table l). The composition is consistent with peptide D extending from position 304 to the tryptic cleavage site at Arglm. The only cysteine in this peptide is Cysym. We conclude that K; aerogenes urease possesses a 160 Table 1 Amino Acid Composition of the Essential Cysteine-Containing Peptide Amino Predicted Experimental Acid Asx Glx Ser Gly His Arg Thr Ala Pro Cys1 Tyr Val Met Val Ile Leu Phe Lys Trp \IKO OHpNOONOOHWNHNNHb-itb :3 Q0 0 o o NprpomH(nmpp(pr-Jp\lii. OOHQNONNI—‘l—‘HWNHWOHWQ 1 Quantitated as carboxymethyl cysteine (although modifica- tion reactions were carried out using IAM, the carboxam- idomethyl derivative is hydrolyzed to the carboxymethyl derivative during acid hydrolysis in 6 N HCl). 2 not detected 161 single essential cysteine residue, and identify this residue as Cyslm. Comparison of K. aerogenes urease sequence to other ureases- Ureases from K. aerogenes (ll), jack bean (17), Helicobacter pylori (4, 6), Ureaplasma urealyticum (3), Proteus mirabilis (5), and Proteus vulgaris (10) have been sequenced. Only one cysteine (Cysxm by K; aerogenes num- bering or Cysan by jack bean numbering) is conserved among all species. As shown in Figure 7, the sequence immediately surrounding the essential cysteine, is identical at 22 of 29 amino acids. This region also contains three conserved his— tidine residues at 312, 320 and 321; histidines have been suggested to participate in binding the active site nickel ions (14, 7). Additionally, one of these histidines could interact with the essential thiol to enhance its reactivity at low pH values (20). Although data from both the bac- terial and plant urease studies are consistent in iden— tifying the essential cysteine, its precise role in catalysis remains unknown. 162 ** *‘k'k'k'k * * *‘k ** *** ****** *‘k ** ** Jb 560-VCGIKNVLPS STNPTRPLTS NTIDEHLDML MVCHHLDREI PEDLAFAHSR-610 Hp 289-VAGEHNILPA STNPTIPFTV NTEAEHMDML MVCHHLDKSI KEDVQFADSR-339 Uu 293-SVKYAHILPA STNPTIPYTV NTIAEHLDML MVCHHLNPKV PEDVAFADSR-343 PV 287—SVGEPNILPA STNPTMPYTI NTVDEHLDML MVCHHLDPSI PEDVAFAESR-337 Pm 287-SVGEPNILPA STNPTMPYTI NTVDEHLDML MVCHHLDPSI PEDVAFADSR-337 Ka 287-ACAHPNILPS STNPTLPYTL NTIDEKLDML MVCHHLDPDI AEDVAFAESR-337 Figure 7. Comparison of urease sequences surrounding the essential cysteine residue. The two letter abbrevia- tions stand for: Jb-jack bean (Takishima et al., 1988); Hp-Helicobacter pylori (Labigne et al., 1991); Uu-Ureaplasma urealyticum (A. Blanchard, 1990); Pv-Proteus vulgaris (Morsdorf & Kaltwasser, 1990); Pm-Proteus mirabilis (Jones & Mobley, 1989); Ka-Klebsiella aerogenes (Mulrooney & Hausinger, 1990). The indicated amino acid positions refer to the a-subunits of the bacterial ureases. Residues which are identical among all sequences are marked by an asterisk; the conserved cysteine and the three conserved histidines are highlighted in bold. 10. ll. 12. 13. 14. 15. REFERENCES Andrews, R. K., Blakeley, R. L. & Zerner. B. (1988) in The Bioinorganic Chemistry of Nickel (Lancaster, J. R., Jr ed) pp. 141—165, VCH Publishers, Inc., New York Allen, G. (1981) in Laboratory Techniques in Biochemistry and Mblecular Biology. (T. S. Work & R. H. Burdon, eds) pp. 52-58, North Holland Pub. Co., NY. Blanchard, A. (1990) Molec. Microbiol. 4, 669—676 Clayton, C. L., Pallen, M. J., Kleanthous, H., Wren, B. W., & Tabaqchali, S. (1990) NUcleic Acids Res. 18, 362 Jones, B. D. & Mobley, H. L. T. (1989) J. Bacteriol. 171, 6414-6422 Labigne, A., Cussac, V. & Courcoux, P. (1991) J. Bacteriol. 173, 1920-1931 Lee, M. H., Mulrooney, S. B. & Hausinger, R. P. (1990) J. Bacteriol. 172, 4427-4431 Lowry, O. H., Rosebrough, N. J., Farr, A. L. & Randall, R. J. (1951) J. Biol. Chem. 193, 265-275 Mobley, H. L. T., and Hausinger, R. P. (1989) Microbiol. Reviews 53. 85-108. Marsdorf, G. & Kaltwasser, H. (1990) FEMS.Micro. Let. 66, 67-74 Mulrooney, S. B. & Hausinger, R. P. (1990) J. Bacteriol. 172, 5837-5843 Mulrooney, S. B., Pankratz, H. S. & Hausinger, R. P. (1989) J. Gen Microbiol. 135. 1769-1766 Norris, R. & Brocklehurst, K. (1976) Biochem. J. 159, 245—257. Sakaguchi, K., Mitsui, K., Kobashi, K., & Hase, J. (1983) J. Biochem. 93. 681-686 Sakaguchi, K., Mitsui, K., Nakai, N. & Kobashi, K. (1984) J. Biochem. 96. 73—79 163 l6. 17. 18. 19. 20. 21. 164 Takishima, K., Mamiya, G. & Hata, M. (1983) in Frontiers in Biochemical and Biophysical Studies of Proteins and Membranes (T. Y. Liu, S. Sakakibara, A. N. Schecter, K. Yagi, H. Yajima, and K. T. Yasunobu, eds.) pp. 193—201, Elsevier, N.Y. Takishima, K,., Suga, T. & Mamiya, G. (1988) Eur. J. Biochem. 175. 151-165 Todd, M. J. & Hausinger, R. P. (1987) J. Biol. Chem. 262. 5963-5967 Todd, M. J. & Hausinger, R. P. (1989) J. Biol. Chem. 264. 15835-15842 Todd, M. J. & Hausinger, R. P. (1991) J. Biol. Chem. (in press). Winkler, R. G., Blevins, D. G., Polacco, J. C. & Randall, D. D. (1988) Trends Biochem. Sci. 13. 97-100 SUMMARY and FUTURE DIRECTIONS Upon initiation of this project little was known about microbial ureases. Most research on this medically and ag- riculturally important enzyme was done by scientists studying mixed microbial populations, whole cells, or impure proteins. The few known characteristics of microbial ureases were similar to the extensively studied plant enzyme, thus some researchers exploited the plant enzyme to draw conclusions about bacterial ureases! This chapter summarizes the purification and characterization of urease from K. aerogenes, describes active site studies using competitive inhibitors and amino acid modifying reagents, and exposes interesting questions arising from this work. Urease purification and characterization- Urease of the Gram negative enteric bacterium, K. aerogenes, was purified to homogeneity and was demonstrated to have an a2B4‘y4 subunit structure, containing 4 nickel (chapter 2) . This was not expected since the well-studied plant enzyme has a homo-hexameric structure containing 12 nickel dispersed among 6 active sites. Using the tight-binding inhibitor PPD we demonstrated that K. aerogenes urease has two active sites (Chapter 3); therefore both enzymes have 2 165 166 mol nickel/mol active unit. The “25474 subunit stoichio— metry has not been definitively determined since quan- titation assumes each subunit has similar binding affinity for dye; however the presence of three distinct subunits was confirmed through the DNA sequencing efforts of Scott Mulrooney. The three subunits share significant homology to the jack bean enzyme and are consistent with gene-fusion in the latter case. Gel filtration chromatography in the presence of salts, glycols, DMSO, reductants, and detergents failed to cause subunit dissociation (Chapter 2). Perhaps moderate concentrations of chaotropic reagents (ag.2.0 M guanidine-HCl) will be found to disrupt subunit interactions, allowing investigation of the forces determining quaternary structure and identification of the nickel binding subunit(s). Not only are the subunits held together tightly, but the nickel is also tightly bound. Nickel is released at pH values below 4.5 and examination of the holoenzyme treated at low pH may reveal increased reactivity of the nickel ligands. Low pH treatment may also lead to replacement of one or both of the nickel with other metals, allowing further spectroscopic and biophysical characterization. Nothing, however, is known of the stability of nickel- depleted urease apo-enzyme to low pH; a low-pH treatment could "irreversibly" denature the protein. Purified urease can migrate as several distinct, active isozymes through native gel electrophoresis. SDS PAGE of 167 2.0— . r4000 [ A 7 “ *4 d L5~ F i A « 1 T.,, n E .1 “ .4; E '-3 “IE, \ ._ o H o- >\ s. U l N g 1.0— 9 ~2ooo < E g .9 E ~2‘\ .E r: _. a Q) 0 O x +1 0 E U a. as— L m _1 “ L 0.0 i r 0 b0 (o— 14 Fraction Figure 1. Separation of urease isozymes on Mono-Q. Urease was applied to a 1.0 x 10 cm Mono-Q equilibrated with 1 mM EDTA, 20 mM phosphate, pH 6.5, containing 1 mM BtME. Urease protein was eluted with an increasing gradient of KCl. Protein, activity, and nickel were determined as described in Chapter 2. 168 each active band showed identical subunit stoichiometry, as assessed by scanning densitometry. These isozymes can also be separated by either ion exchange chromatography or gradient hydrophobic chromatography. The major active form (isozyme B) migrates faster under native gel electrophoresis and elutes later from anion exchange chromatography, consistent with a greater anionic character. One possible explanation for the multiple forms is that the minor form has less than 4 mol nickel/mol enzyme. Figure l is a preliminary experiment determining specific activity, protein, and nickel content of urease eluting from Mono-Q anion exchange chromatography. The isozyme which elutes first (A) has only 75% of the nickel (and 75% of the specific activity) of the isozyme which elutes later (B) consistent with A having less than stoichiometric amounts of nickel. While isozyme B is very stable, it will (over time) form some isozyme A; isozyme A looses activity quickly at low pH and has never been observed to form isozyme B. To simplify experiments, all inhibitor and active site studies in chapters 3-5 (and all biophysical characterizations) employed the second isozyme only. Interpretation of urease structural studies would be easier with knowledge of the urease crystal structure. Methods to crystallize plant ureases have been known for 65 years, yet poor crystal quality and large unit cell have precluded solving the crystal structure. Crystals of the K. aerogenes enzyme have been obtained (A. Karplus, in 169 progress) and structural studies are continuing. Integration of crystal structure with chemical and genetic studies should contribute to understanding the mechanism of urea hydrolysis. Urease inhibitors— The intelligent design of useful urease inhibitors requires detailed knowledge of enzyme— inhibitor interactions. AHA and PPD were found to be slow binding competitive inhibitors of K. aerogenes urease (Chapter 3). A slowly processed substrate is kinetically indistinguishable from a slow-binding inhibitor, therefore assays for PPD hydrolysis should be done. If PPD is hydro— lyzed, the rate of hydrolysis may correspond to the rate of formation of enzyme-inhibitor complex, or to the rate of enzyme reactivation. A thorough examination of the pH— dependence for PPD inhibition of urease may help elucidate the mechanism of urea hydrolysis. Preliminary studies using AHA-urease indicated that the rate of reactivation (My increased below pH 6.5, perhaps indicating the enzyme pl; of 6.55 (required to be deprotonated for urea catalysis) must be deprotonated for the formation of a tight binding transition state complex with AHA or PPD. With the advent of molecular biology and the develop- ment of expression vectors producing large quantities of protein, biophysical and spectroscopic studies on the urease nickel center are now possible. Magnetic susceptibility and MCD measurements on urease are consistent with two nickel 170 per active site being antiferromagnetically coupled in the presence of B—ME. Similar studies using the slow-binding inhibitors PPD and AHA may confirm the proposed bridging interactions of these competitive inhibitors with urease. Additional evidence linking the interactions of AHA and PPD with the nickel center may be deduced if the UV-visible spectrum of K. aerogenes urease changes upon binding of these inhibitors. The rate of spectral change may correspond to the rapid equilibration between enzyme and AHA or the slow step leading to our proposed "bridging" intermediate. Urease active site amino acids- Ureases from all sources examined are susceptible to inactivation by thiol specific reagents. Competitive inhibitors modified the reactivity of an essential thiol in K. aerogenes urease and protected one thiol per catalytic unit from modification (Chapter 4). The essential cysteine of the jack bean enzyme is thought to act as a general acid, donating a hydrogen to the ammonia molecule as it leaves. We have demonstrated that at catalytically active pH values, this essential cysteine has appreciable nucleophilic character and that it interacts with another ionizable enzyme group (Chapter 4). The identity of this second ionizable group is not known, however the model of Zerner predicts three additional ion- izable groups in close proximity to the active site thiol: 171 a carboxylate, an un—identified base, and a nickel—bound hydroxide. . Only indirect evidence for a carboxyl group at the active site of urease has been published: comparison of R; values for various nickel binding inhibitors (Chapter 3) and (app values for several thiol modifying reagents (Chapter 4) were consistent with repulsion of negatively charged molecules by the urease active site. Preliminary studies of K. aerogenes urease showed no loss of activity upon exposure to Woodwards Reagent K, at 37'C. However this reagent decomposes rapidly. More thorough analysis using carbo— diimides or trialkyloxonium salts are obviously needed. Prior protection of the essential thiol with a reversible modifying reagent (ag.nmdification by disulfide reagents is reversible) may be necessary. Other amino acid specific reagents may be used to identify amino acids which participate in catalytic turnover. For example, DEP has been demonstrated to quickly inactivate urease holoenzyme (Mann Hyung Lee), consistent with a histidine being essential for activity; however the effect of active site ligands and the identity of the amino acid modified is not known. Tetranitromethane could be used to modify tyrosine side chains; lysine could be guanidinated using O-methyl isourea or reductively methylated with formate/NaCNBH4. Preliminary experiments using the latter two lysine modifying reagents were hampered because the reagents interfered with the indophenol assay. Figure 2 172 U 5' 6’ c o 4_e ( . Control . . ' D _2_ . . A . o 2.4-pentonedlone >° . \_ 1 3 _4_ E ° 2,3—butonedione _6_. phenylglyoxol T T T r r I I T T r T T T T r T T ¢ T r T T T T I I I I O 25 50 75 100 125 Thne (nfinutes) Figure 2. Pseudo-first order inactivation of urease by arginine modifying reagents. Urease was incubated at 37'C in 8 mM EDTA, 80 mM HEPES, pH 7.75 alone, or in the presence of 200 mM 2,4—pentane dione, 50 mM 2,3-butane dione, or 100 mM phenylglyoxal. At the indicated intervals, a small aliquot was diluted into the standard assay mixture to determine activity. 173 demonstrates the pseudo—first order loss of K. aerogenes urease activity by three arginine modifying reagents: 2,4— pentanedione, 2,3-butanedione, and phenylglyoxal. A comprehensive analysis of the effects of arginine modifying reagents will include determination of: the apparent second-order rate constants (and the effect of pH on these rate constants), the effects of active site ligands, and the effect of prior modification of the essential thiol. Identification of the essential arginine residue would naturally follow. Clearly, expanded investigation of essential amino acids could contribute to our understanding of the enzyme mechanism” Once an amino acid is identified (as Cys319 was identified as essential for K. aerogenes urease activity; Chapter 5), site directed mutagenesis, (varying amino acid size, charge, polarity, or ability to form hydrogen bonds) could distinguish the properties of the amino acid involved in catalysis. Similar site directed mutagenesis studies could be targeted towards proposed nickel ligands. Simultaneous biophysical, spectroscopic, chemical, kinetic, and genetic analysis of the urease active site should generate complementary data, creating a thorough understanding of the catalytic mechanism and leading to the rational design of potent inhibitors for this medically and agriculturally important enzyme. APPENDIX SATURATION MAGNETIZATION OF THE NICKEL CENTERS IN UREASE. E,P, Day, J. Peterson+, M.J. Todd*, and R.P. Hausinger* University of Minnesota, Minneapolis, MN 55455. +University of Alabama, Tuscaloosa, AL 35487. *Michigan State University, East Lansing, MI 48824. The two nickel ions at the active site of urease are believed to be in the Ni(II) oxidation state. We have measured the saturation magnetization of urease isolated from Klebsiella aerogenes to determine its spin, spin concentration, average g value, and zero-field splitting. Activity was measured before and after the magnetization data were collected. EPR was used to quantitate Fe(III) impurities and metal analysis to quantitate the nickel. Spin S=l paramagnetism accounted for roughly half the nickel implying that half the nickel is low spin (8:0). No evidence for exchange coupling between the nickel ions was found. 174